the regulation of tgf signal transduction · signal transduction in metazoans and emphasize events...

16
3699 Transforming growth factor (TGF) pathways are implicated in metazoan development, adult homeostasis and disease. TGF ligands signal via receptor serine/threonine kinases that phosphorylate, and activate, intracellular Smad effectors as well as other signaling proteins. Oligomeric Smad complexes associate with chromatin and regulate transcription, defining the biological response of a cell to TGF family members. Signaling is modulated by negative-feedback regulation via inhibitory Smads. We review here the mechanisms of TGF signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor (TGF) family consists of 33 members, most of which encode dimeric, secreted polypeptides that control developmental processes, ranging from gastrulation and body axis asymmetry to organ-specific morphogenesis and adult tissue homeostasis (reviewed by Derynck and Miyazono, 2008). In addition to TGFs, this family includes the bone morphogenetic proteins (BMPs), growth and differentiation factors (GDFs), activins and nodal. The TGF family is conserved throughout metazoan evolution. At the cellular level, TGF family members regulate cell growth, differentiation, adhesion, migration and death, in a developmental context-dependent and cell type-specific manner. For example, TGF more often inhibits, but sometimes also stimulates, cell proliferation (reviewed by Yang and Moses, 2008). Furthermore, nodal signaling sometimes inhibits, whereas BMP promotes, cell differentiation, as in stem cells (Watabe and Miyazono, 2009). As TGF ligands act multifunctionally in numerous tissue types, they also play complex roles in various human diseases, ranging from autoimmune to cardiovascular diseases and cancer (reviewed by Gordon and Blobe, 2008; Massagué, 2008). Here we review the core components of the TGF family and their signaling engines, as part of a Minifocus in this issue on TGF signaling (see Box 1), and discuss emerging concepts concerning the regulatory mechanisms of TGF pathways at the receptor, cytoplasmic and nuclear level. We also highlight recent discoveries that are of particular developmental relevance. The TGF family The development of the axes and the asymmetry of the animal body depends on the localized action of extracellular signals, such as the Wnt, nodal and BMP ligands. Gradients of these ligands, their extracellular regulators and the competence of receptors in responding cells, play important roles during tissue morphogenesis (Affolter and Basler, 2007; Smith and Gurdon, 2004). TGF family members also contribute to tissue patterning and are important regulators of stem cell self-renewal and differentiation (see Box 2) (De Robertis and Kuroda, 2004; Watabe and Miyazono, 2009). The TGF morphogens include numerous secreted and conserved polypeptides (Table 1), which emerged at the onset of multicellular (metazoan) life (Huminiecki et al., 2009). Structurally, this family is characterized by a specific three-dimensional fold and by a conserved number and spacing of cysteine residues in the C- terminus of the mature polypeptide (Derynck and Miyazono, 2008). The prototypic TGF isoforms (TGF1, 2, 3), and the related inhibin polypeptides that make up the activin and inhibin members, have nine characteristic cysteines, eight of which form four intramolecular disulfide bridges, while one intermolecular bond links the two monomers. The inhibin polypeptides, BMPs and GDFs have seven cysteines, of which six form intramolecular and one intermolecular bridges. The lefty proteins, GDF3, GDF9 and BMP15A have six cysteines in their mature sequence and lack the intermolecular bridge between the two monomers. The lack of covalent dimers provides regulatory flexibility; for example, lefty forms non-covalent complexes with nodal and binds to the glycosylphosphatidylinositol (GPI)-anchored co-receptor of the epidermal growth factor-Cripto/FRL-1/Cryptic (EGF-CFC) family, leading to the inhibition of nodal signaling (Chen and Shen, 2004). Xenopus laevis expresses TGFs, nodal, activins, BMPs and GDFs (De Robertis and Kuroda, 2004), and additional unique family members, such as the mesoderm-inducer Derrière, and the six nodal-related proteins XNR1-6 (Eimon and Harland, 2002; Onuma et al., 2002; Ramis et al., 2007). Drosophila melanogaster has only seven TGF family members (Table 1). The BMP-like ligands Decapentaplegic (Dpp) and Screw (Scw) regulate dorsoventral pattering and the differentiation of imaginal discs, such as the wing disc (Affolter and Basler, 2007; Serpe et al., 2005). The BMP-like Glass bottom boat (Gbb) regulates brain and wing disc differentiation (Bangi and Wharton, 2006; Goold and Davis, 2007). The activin-like dActivin (Act – FlyBase) and Dawdle (Daw) ligands have tissue-specific roles (for example, in the larval brain), whereas much remains to be understood about the functions of Maverick, the GDF8 (myostatin)-like ligand, and of Myoglianin, which are expressed in endodermal and mesodermal cells (Lee- Hoeflich et al., 2005; Nguyen et al., 2000; Zhu et al., 2008). In Caenorhabditis elegans, the BMP-like DBL-1, and the TGF-like DAF-7, regulate body length and the dauer pathway, a special environmental adaptation of earthworms, respectively (Table 1) (reviewed by Savage-Dunn, 2005). The other three ligands, TIG-2, TIG-3 and UNC-129, are as yet unexplored. Development 136, 3699-3714 (2009) doi:10.1242/dev.030338 The regulation of TGF signal transduction Aristidis Moustakas and Carl-Henrik Heldin Ludwig Institute for Cancer Research, Uppsala University, BMC, Box 595, SE-75124 Uppsala, Sweden. [email protected]; [email protected] REVIEW Box 1. Minifocus on TGF signaling This article is part of a Minifocus on TGF signaling. For further reading, please see the accompanying articles in this collection: ‘The extracellular regulation of bone morphogenetic protein signaling’ by David Umulis, Michael O’Connor and Seth Blair (Umulis et al., 2009); ‘Informatics approaches to understanding TGF pathway regulation’ by Pascal Kahlem and Stuart Newfeld (Kahlem and Newfeld, 2009); and ‘TGF family signaling: novel insights in development and disease’, a review of a recent FASEB Summer Conference on TGF signaling by Kristi Wharton and Rik Derynck (Wharton and Derynck, 2009). DEVELOPMENT

Upload: others

Post on 03-Aug-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3699

Transforming growth factor (TGF) pathways are implicated inmetazoan development, adult homeostasis and disease. TGFligands signal via receptor serine/threonine kinases thatphosphorylate, and activate, intracellular Smad effectors as wellas other signaling proteins. Oligomeric Smad complexesassociate with chromatin and regulate transcription, definingthe biological response of a cell to TGF family members.Signaling is modulated by negative-feedback regulation viainhibitory Smads. We review here the mechanisms of TGFsignal transduction in metazoans and emphasize events crucialfor embryonic development.

IntroductionThe human transforming growth factor (TGF) family consists of33 members, most of which encode dimeric, secreted polypeptidesthat control developmental processes, ranging from gastrulation andbody axis asymmetry to organ-specific morphogenesis and adulttissue homeostasis (reviewed by Derynck and Miyazono, 2008). Inaddition to TGFs, this family includes the bone morphogeneticproteins (BMPs), growth and differentiation factors (GDFs), activinsand nodal. The TGF family is conserved throughout metazoanevolution. At the cellular level, TGF family members regulate cellgrowth, differentiation, adhesion, migration and death, in adevelopmental context-dependent and cell type-specific manner. Forexample, TGF more often inhibits, but sometimes also stimulates,cell proliferation (reviewed by Yang and Moses, 2008). Furthermore,nodal signaling sometimes inhibits, whereas BMP promotes, celldifferentiation, as in stem cells (Watabe and Miyazono, 2009). AsTGF ligands act multifunctionally in numerous tissue types, theyalso play complex roles in various human diseases, ranging fromautoimmune to cardiovascular diseases and cancer (reviewed byGordon and Blobe, 2008; Massagué, 2008).

Here we review the core components of the TGF family and theirsignaling engines, as part of a Minifocus in this issue on TGFsignaling (see Box 1), and discuss emerging concepts concerning theregulatory mechanisms of TGF pathways at the receptor,cytoplasmic and nuclear level. We also highlight recent discoveriesthat are of particular developmental relevance.

The TGF familyThe development of the axes and the asymmetry of the animal bodydepends on the localized action of extracellular signals, such as theWnt, nodal and BMP ligands. Gradients of these ligands, theirextracellular regulators and the competence of receptors inresponding cells, play important roles during tissue morphogenesis(Affolter and Basler, 2007; Smith and Gurdon, 2004). TGF familymembers also contribute to tissue patterning and are importantregulators of stem cell self-renewal and differentiation (see Box 2)(De Robertis and Kuroda, 2004; Watabe and Miyazono, 2009).

The TGFmorphogens include numerous secreted and conservedpolypeptides (Table 1), which emerged at the onset of multicellular(metazoan) life (Huminiecki et al., 2009). Structurally, this familyis characterized by a specific three-dimensional fold and by aconserved number and spacing of cysteine residues in the C-terminus of the mature polypeptide (Derynck and Miyazono, 2008).The prototypic TGF isoforms (TGF1, 2, 3), and the relatedinhibin polypeptides that make up the activin and inhibinmembers, have nine characteristic cysteines, eight of which formfour intramolecular disulfide bridges, while one intermolecular bondlinks the two monomers. The inhibin polypeptides, BMPs andGDFs have seven cysteines, of which six form intramolecular andone intermolecular bridges. The lefty proteins, GDF3, GDF9 andBMP15A have six cysteines in their mature sequence and lack theintermolecular bridge between the two monomers. The lack ofcovalent dimers provides regulatory flexibility; for example, leftyforms non-covalent complexes with nodal and binds to theglycosylphosphatidylinositol (GPI)-anchored co-receptor of theepidermal growth factor-Cripto/FRL-1/Cryptic (EGF-CFC) family,leading to the inhibition of nodal signaling (Chen and Shen, 2004).

Xenopus laevis expresses TGFs, nodal, activins, BMPs andGDFs (De Robertis and Kuroda, 2004), and additional uniquefamily members, such as the mesoderm-inducer Derrière, and thesix nodal-related proteins XNR1-6 (Eimon and Harland, 2002;Onuma et al., 2002; Ramis et al., 2007). Drosophila melanogasterhas only seven TGF family members (Table 1). The BMP-likeligands Decapentaplegic (Dpp) and Screw (Scw) regulatedorsoventral pattering and the differentiation of imaginal discs, suchas the wing disc (Affolter and Basler, 2007; Serpe et al., 2005). TheBMP-like Glass bottom boat (Gbb) regulates brain and wing discdifferentiation (Bangi and Wharton, 2006; Goold and Davis, 2007).The activin-like dActivin (Act – FlyBase) and Dawdle (Daw)ligands have tissue-specific roles (for example, in the larval brain),whereas much remains to be understood about the functions ofMaverick, the GDF8 (myostatin)-like ligand, and of Myoglianin,which are expressed in endodermal and mesodermal cells (Lee-Hoeflich et al., 2005; Nguyen et al., 2000; Zhu et al., 2008). InCaenorhabditis elegans, the BMP-like DBL-1, and the TGF-likeDAF-7, regulate body length and the dauer pathway, a specialenvironmental adaptation of earthworms, respectively (Table 1)(reviewed by Savage-Dunn, 2005). The other three ligands, TIG-2,TIG-3 and UNC-129, are as yet unexplored.

Development 136, 3699-3714 (2009) doi:10.1242/dev.030338

The regulation of TGF signal transductionAristidis Moustakas and Carl-Henrik Heldin

Ludwig Institute for Cancer Research, Uppsala University, BMC, Box 595, SE-75124Uppsala, Sweden.

[email protected]; [email protected]

REVIEW

Box 1. Minifocus on TGF signalingThis article is part of a Minifocus on TGF signaling. For further reading,please see the accompanying articles in this collection: ‘Theextracellular regulation of bone morphogenetic protein signaling’ byDavid Umulis, Michael O’Connor and Seth Blair (Umulis et al., 2009);‘Informatics approaches to understanding TGF pathway regulation’by Pascal Kahlem and Stuart Newfeld (Kahlem and Newfeld, 2009);and ‘TGF family signaling: novel insights in development and disease’,a review of a recent FASEB Summer Conference on TGF signaling byKristi Wharton and Rik Derynck (Wharton and Derynck, 2009).

DEVELO

PMENT

Page 2: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3700

Although specification of body asymmetry is a fundamentalfunction of TGF-like proteins during early embryogenesis, theidentification of genes that encode a complete TGF pathway in theprimitive metazoan Trichoplax adhaerens, a two-cell-layered animalthat lacks obvious body asymmetry, suggests that these morphogensmight have played a fundamental role in the specification of themulticellularity that precedes body asymmetry during animalevolution (Huminiecki et al., 2009).

TGF secretion and extracellular regulationAll TGF ligands are synthesized as precursor proteins with a longerN-terminal pro-peptide followed by a shorter C-terminal maturepolypeptide (reviewed by ten Dijke and Arthur, 2007). Intermoleculardisulfide linkages pair dimers of these precursors via conservedcysteine residues in the pro-peptide and mature peptide sequence.While precursor proteins are in the secretory pathway, furin-likeproteases cleave the pro-peptide from the mature peptide. The TGFpro-peptide, called the latency-associated peptide (LAP), continues toscaffold the smaller mature peptide within its core, serving as achaperone during exocytosis of the complex. It also mediates thedeposition of TGF in the extracellular matrix (ECM) through itscovalent association with large secreted proteins called latent TGF-binding proteins (LTBPs), and with ECM proteins, such as fibronectinand fibrillin 1 (reviewed by Rifkin, 2005). Activation of the mature C-terminal dimeric ligands from their matrix-deposited, multi-protein‘cages’ relies on several proteases, including elastase (which cleavesfibrillin 1), BMP1/Tolloid family proteases (which cleave LTBPs),and matrix metalloproteases, such as MMP2 (which cleave TGFLAPs) (reviewed by ten Dijke and Arthur, 2007).

The ability of the TGF LAP to maintain the ligand in an inactivestate is conserved among some TGF family ligands, such as GDF8and GDF11 (Ge et al., 2005; Wolfman et al., 2003). However, thenodal, BMP4 and BMP7 pro-peptides do not act as extracellular

antagonists, but instead regulate mature ligand stability andprocessing, including ligand degradation in lysosomes, which limitsligand availability (Degnin et al., 2004; Dick et al., 2000; Le Goodet al., 2005). Similarly, the nodal pro-peptide associates with itsEGF-CFC family co-receptor Cripto in secretory vesicles near thecell surface (Blanchet et al., 2008). Cripto also forms complexeswith mature nodal and enhances signaling via the receptor kinasecomplex (see below) (Bianco et al., 2004). Recent evidencedemonstrates that the signaling Cripto-nodal-receptor complexenters a specialized endocytic pathway that is characterized by theprotein flotillin, possibly en route to its final degradation.Interestingly, many other TGF ligands are inactivated in theextracellular space by antagonists, such as noggin and chordin,which inhibit BMPs, and follistatin, which inhibits activins(Gazzerro and Canalis, 2006; Harrison et al., 2005). Theseextracellular antagonists help to establish the morphogen gradientsthat pattern early embryos, as discussed in an accompanying review(Umulis et al., 2009) (see Box 1).

The TGF receptor familyAll TGF ligands transmit biological information to cells by bindingto type I and type II receptors that form heterotetrameric complexesin the presence of the dimeric ligand (reviewed by Wrana et al.,2008). Five type II and seven type I receptors exist in humans andother mammals, and are characterized by a cytoplasmic kinasedomain that has strong serine/threonine kinase activity and weakertyrosine kinase activity, which classifies them as being dual-specificity kinases (Table 1) (reviewed by ten Dijke and Heldin,2006). The type I receptors are also known as activin receptor-likekinases (ALKs), a nomenclature that is employed to tackle theproblem of one ligand signaling via many receptors, or many ligandssignaling via the same receptor. TGF ligands also interact with co-receptors that either facilitate or limit receptor kinase signaling. Inaddition to the EGF-CFC/Cripto co-receptors discussed above, typeIII receptors, such as endoglin and the proteoglycan betaglycan(TGFR3; TRIII), regulate TGF signaling in mammals, as doesthe repulsive guidance molecule (RGM, also known as Dragon)family of co-receptors (reviewed by Wrana et al., 2008) (Table 1).

Ligand binding links the constitutively active type II receptorkinases to the dormant type I receptor kinases, allowing the type IIreceptor to phosphorylate the juxtamembrane part of thecytoplasmic domain of the type I receptor (Fig. 1), turning onreceptor kinase activity (reviewed by Wrana et al., 2008). Recentstructural analysis of TGF and BMP ligands bound to theirrespective type I and type II receptor ectodomains shows that TGFligands contact both receptors tightly, whereas the evolutionarilymore ancient BMPs associate more loosely with their receptors(Groppe et al., 2008). Binding of TGF to TRII (TGFR2) createsthe interface required for TRI (ALK5; TGFR1) type I receptorrecruitment to the complex.

D. melanogaster has five TGF family receptors, including thetype II receptors Punt (Put) and Wishful thinking (Wit), which bindthe BMP-like ligands Dpp, Gbb and Scw during fly developmentand which form complexes with the type I receptors Thickveins(Tkv) and Saxophone (Sax) (Table 1) (Affolter and Basler, 2007;Goold and Davis, 2007; Serpe et al., 2005). Put and Wit also pairwith the type I receptor Baboon (Babo) to mediate activin-likesignals from dActivin and Daw (Zhu et al., 2008). Theaccompanying review by Umulis et al. (Umulis et al., 2009)discusses how, in the developing wing disc, a gradient of BMP-likesignaling activity is achieved by the dual contribution of Dpp andGbb, which differentially bind to distinct receptor complexes.

REVIEW Development 136 (22)

Box 2. Role of TGF/BMP signaling in embryonic stemcellsStem cells exhibit self-renewing capacity and pluripotency ingenerating the multitude of embryonic and adult cell types of themetazoan body (reviewed by Rossi et al., 2008). Growth factors, suchas TGF and FGF, regulate stem cell self-renewal and differentiation.FGF2, the most widely used growth factor that supports mouse andhuman embryonic stem cell (ESC) self-renewal in culture, inducesTGF/activin ligands and receptors while suppressing BMP-likeactivities (Greber et al., 2007; Ogawa et al., 2007). Furthermore,pharmacological inhibitors of the TGF/nodal type I receptor familysuppress human and mouse ESC self-renewal (Ogawa et al., 2007).In general, TGF inhibits differentiation of pluripotent progenitorcells, whereas BMP induces their differentiation (Watabe andMiyazono, 2009) (Fig. 7A,B).

To promote self-renewal of ESCs, TGF/nodal signaling activatesSMAD2 and SMAD3, which directly induce Nanog, one of the crucialstem cell transcription factors (Xu, R. H. et al., 2008). TGF and FGFsignaling synergize by enhancing binding of Smad complexes to theNanog promoter. Interestingly, NANOG provides a molecular link forthe antagonism between TGF (the self-renewing factor) and BMP(the differentiation factor) in ESCs. NANOG binds to SMAD1,inhibiting its transcriptional activity and limiting the BMP signalingpotential that promotes early mesodermal differentiation or tissue-specific differentiation later in development (Suzuki et al., 2006). Thisexample is likely to be expanded to additional regulators of ESC self-renewal and differentiation as a result of genome-wide screens forthe transcription and signaling factors of these pathways (Chen etal., 2008).

DEVELO

PMENT

Page 3: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3701REVIEWDevelopment 136 (22)

C. elegans has three TGF family receptors (Table 1) (Pattersonand Padgett, 2000). In the dauer pathway, DAF-7 signals via the typeII receptor DAF-4 and the type I receptor DAF-1. In the Sma/Mabpathway, which regulates body length, tail development and innateimmunity, DBL-1 signals via DAF-4 and the SMA-6 type I receptor.Which of these receptors mediate signals by UNC-129, TIG-2 andTIG-3 remains unknown.

Finally, T. adhaerens has one type II and three type I receptors(Huminiecki et al., 2009), which is compatible with a model inwhich the type II receptor represents the ligand-recognizing core,whereas the type I receptor is the downstream signaling effectorthat defines biological responses and that has diverged morerapidly to serve the new developmental processes of morecomplex organisms.

The Smad familyThe activated type I receptor phosphorylates cytoplasmic proteinsof the Smad family in their C-terminal regions (Figs 1 and 2). Smadsconsist of three domains: (1) an N-terminal Mad-homology 1 (MH1)domain that can interact with other proteins and carries nuclearlocalization signals (NLSs) and a DNA-binding domain; (2) amiddle linker domain that interacts with prolyl-isomerases andubiquitin ligases and that is enriched in prolines andphosphorylatable serines or threonines; and (3) a C-terminal MH2domain that binds to type I receptors and can interact with otherproteins, and that mediates Smad homo- and hetero-oligomerizationand mediates the transactivation potential of nuclear Smadcomplexes (Fig. 2) (reviewed by ten Dijke and Heldin, 2006). TheC-terminal phosphorylation of receptor-activated (R) Smads allows

Table 1. TGF pathways in humans, flies and worms

H. sapiens

Pathway BMP GDF Activin TGF AMH Inhibitors

Ligand BMP2, 4BMP5, 6, 7BMP8A, 8BBMP9, 10

GDF5, 6, 7GDF9b

GDF10, 11GDF15 (MIC1)------------------

GDF1, 3GDF8 (MYO)

GDF9

Inhibin AInhibin B

Nodal

TGF 1TGF 2TGF 3

AMH (MIS) BMP3Inhibin Inhibin CInhibin ELEFTYALEFTYB

RII BMPRIIActRIIA, ActRIIB

BMPRIIActRIIA, ActRIIB

ActRIIAActRIIB

T RII AMHRII N/A

RI BMPRIA (ALK3)BMPRIB (ALK6)

ALK2ALK1

BMPRIA (ALK3)BMPRIB (ALK6)

ALK2-------------------ActRIB (ALK4)

ALK7T RI (ALK5)

ActRIB (ALK4)ALK7

T RI (ALK5)--------------

ALK1ALK2

BMPRIA (ALK3)

BMPRIA (ALK3)BMPRIB (ALK6)

ALK2

N/A

RIII RGMa, b, c (+) Cripto 3 (+) Cripto 3 (–)Cripto 1 (+)

T RIII (+)Endoglin (+)Cripto 3 (–)

? T RIII (–)Cripto 3 (–)

R-Smad SMAD1, 5, 8 SMAD1, 5, 8–-----------------

SMAD2, 3

SMAD2, 3 SMAD2, 3------------------SMAD1, 5, 8

SMAD1, 5, 8 N/A

Co-Smad SMAD4 SMAD4 SMAD4 SMAD4 SMAD4 N/A

I-Smad SMAD6, 7 SMAD6, 7 SMAD7 SMAD7 SMAD6, 7 N/A

D. melanogaster C. elegans

Pathway BMP Activin Other Sma/Mab Dauer

Ligand DppGbbScw

dActivinDaw

MavMyo

DBL-1 DAF-7

RII PutWit

PutWit

? DAF-4 DAF-4

RI TkvSax

Babo ? SMA-6 DAF-1

RIII ? ? ? ? ?

R-Smad Mad dSmad2 ? SMA-2SMA-3

DAF-8DAF-14

Co-Smad Medea Medea ? SMA-4 DAF-3 (?)

I-Smad Dad ? ? TAG-68 (?) TAG-68 (?)Receptors are listed as type II (RII), type I (RI) and type III (RIII) co-receptors. Dashed lines separate groups of ligands or receptors based on the division into BMP andTGF /activin-like pathways. Ligands, type I receptors and R-Smads are color-coded: blue, BMP-like pathways; red, TGF /activin-like pathways. Question marks indicateunassigned signaling relationships. For two human ligands, GDF8 and GDF15, we provide their alternative names [myostatin (MYO) and macrophage inhibitory cytokine 1(MIC1)], as the latter are more commonly used in the literature. LEFTYA and B are also known as LEFTY2 and 1, respectively. The co-receptor T RIII is also known asbetaglycan. Cripto 1 and Cripto 3 are also known as TDGF1 and TDGF3, respectively. In the RIII group (+) or (–) indicates positive or negative effects, respectively, onsignaling by each co-receptor. N/A, not applicable.

DEVELO

PMENT

Page 4: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3702

them to associate with the common-mediator (Co) Smad, SMAD4.The resulting Smad oligomer is thought to consist of a trimer of twoR-Smads and a single SMAD4 (such as a SMAD2-SMAD2-SMAD4 complex, a SMAD3-SMAD3-SMAD4 complex, or aSMAD2-SMAD3-SMAD4 complex), which is then shuttled into thenucleus. Nuclear Smad complexes bind to chromatin, and, togetherwith other transcription factors, regulate target gene expression (Fig.1) (reviewed by Massagué et al., 2005; Schmierer and Hill, 2007).TGF- or BMP-specific Smad complexes induce the expression ofthe inhibitory (I) Smads, SMAD6 and SMAD7 (Figs 1 and 2), whichnegatively regulate signaling strength and duration, thus forming anegative-feedback loop (reviewed by Itoh and ten Dijke, 2007).

All genomes sequenced to date possess the three fundamentalclasses of Smad proteins: R-, Co- and I-Smads (Table 1)(Huminiecki et al., 2009). In D. melanogaster, BMP-like signalingis mediated by a single R-Smad (Mad) and a single Co-Smad(Medea), despite the existence of two type I receptors (Tkv, Sax)(Affolter and Basler, 2007). In the activin-like pathways, the type Ireceptor Babo signals via dSmad2 (Smox) (an R-Smad) and Medea.A single I-Smad (Dad) also operates during wing imaginal discdevelopment (Tsuneizumi et al., 1997). Dad binds and inhibitssignaling from the Dpp/Scw/Gbb type I receptors Tkv and Sax, butnot from the dActivin type I receptor Babo (Kamiya et al., 2008).

In C. elegans, the Sma/Mab pathway engages two R-Smads,SMA-2 and SMA-3, that signal with a single Co-Smad, SMA-4(Patterson and Padgett, 2000). The dauer pathway has two R-Smads,DAF-8 and DAF-14, which have more divergent MH1 domains butconserved MH2 domains that suggest activation by thecorresponding type I receptors. The dauer pathway Co-Smad ispossibly DAF-3, which presents peculiar developmentalcharacteristics. Unlike the Co-Smads in other organisms and SMA-4 in C. elegans, the DAF-3 loss-of-function phenotype is distinctfrom those of loss-of-function mutations in the ligand, receptors orR-Smads of this pathway (Patterson et al., 1997). Furthermore, thereceptors and R-Smad seem to negatively regulate the function ofDAF-3. Thus, DAF-3 is classified as a Co-Smad only on the basisof sequence similarity to other Co-Smads (Huminiecki et al., 2009).Finally, TAG-68 is classified as an I-Smad based on phylogeneticarguments, although functional evidence for such a role is currentlyabsent (Savage-Dunn et al., 2003). T. adhaerens also has threedistinct Smad classes, which suggests that the three differentfunctional features of Smad proteins evolved early during metazoanevolution (Huminiecki et al., 2009).

A fundamental feature of all TGF signaling pathways is theirdivision into TGF-like and BMP-like cascades, a classification basedon the specificity of interaction between the so-called L45 loop of the

REVIEW Development 136 (22)

RI

P

RII

TGFββ

Co-Smad

Smad3

Latent TGFβ

PSmad2

PTight

junction TAK1

JNK

Cytoplasm

Nucleus

RI

P

RII

BMP

Smad1P

Smad5P

RI

P

RII

TGFβ

GSK3βRho

Par6 P

Smurf

Actin

TRAF6

p38

ShcAP

ERK

Co-Smad

Other substrates

Src

Smad7

mRNA

TF

R-Smad PCo-Smad

Chromatin

Smad7

Noggin/BMP

A BC

Smad6/7

Smad8P

mRNA

TFChromatin

P

Rho

?

Actin

Stress fibersFocal adhesionsMotility

ActivinNodal

RI

P

RII

Follistatin/Activin

Smad7

Smad6ROCK

LIMK2

?

ChromatinmRNA

TF

R-Smad PCo-SmadR-Smad

P

R-SmadP

Fig. 1. TGF and BMP signaling. (A,B)The (A) TGF and activin/nodal and (B) BMP pathways, with their corresponding Smad proteins andmechanisms of inhibition by I-Smads (Smad6/7). The latent TGF complex and the extracellular antagonists, follistatin (bound to activin) and noggin(bound to BMP) are shown. (C)Non-Smad signaling pathways downstream of the TGF receptors [RI, TRI (ALK5) and RII, TRII (TGFR2)]. Thenuclear Smad complexes that lead to gene regulation are shown for each pathway. In these complexes, the Smad trimer most likely contains two R-Smad (identical or different) and one Co-Smad subunit. In addition to the major signaling pathways shown, TGF also activates BMP R-Smads incertain contexts (see text). BMP (bone morphogenetic protein), Erk (extracellular signal-regulated kinase), LIMK2 (LIM domain kinase 2), JNK (Jun N-terminal kinase), p38 (p38 MAPK), PAR6 (partitioning-defective 6 homolog), Rho (Ras homolog), ROCK (Rho-associated, coiled-coil-containingprotein kinase), SHCA (SH2 domain-containing sequence A), Smurf (Smad ubiquitylation regulatory factor), Src (Rous sarcoma virus oncoprotein),TGF (transforming growth factor ), TAK1 (TGF-activated kinase 1), TF (transcription factor), TRAF6 (tumor necrosis factor receptor-associatedfactor 6).

DEVELO

PMENT

Page 5: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3703REVIEWDevelopment 136 (22)

type I receptors and the L3 loop of the MH2 domains of R-Smads(Fig. 2) (reviewed by ten Dijke and Heldin, 2006). TGF/activinpathways signal via SMAD2, SMAD3, and BMP/GDF pathways viaSMAD1, SMAD5 and SMAD8. However, in a diversity of cell typesin culture, such as endothelial, immortalized epithelial, adenoma andcarcinoma cell lines and even NIH-3T3 fibroblasts and chondrocytes(Daly et al., 2008; Finnson et al., 2008; Goumans et al., 2003; Liu etal., 2008), TGF signaling can also activate SMAD1 and SMAD5.Originally, TGF was shown to bind to two type I receptors, thusactivating SMAD2 and SMAD3 via TRI, and SMAD1, SMAD5 andSMAD8 via the BMP type 1 receptor ALK1 (ACVRL1). ALK1 isexpressed mainly in endothelial cells, where SMAD2, SMAD3signaling inhibits and SMAD1, SMAD5, SMAD8 signaling promotesproliferation and migration (Goumans et al., 2003). However, TGFcan also activate SMAD1 and SMAD5 via two additionalmechanisms. First, in immortalized EpH4 mouse mammary epithelialcells and in MDA-MB-231 human mammary carcinoma cells, TGFinduces SMAD1 and SMAD5 C-terminal phosphorylation viaheteromeric receptor complexes that form between TRII and TRI,as well as between TRII and the BMP type I receptors ALK2(ACVR1) and BMPRIA (ALK3) (Daly et al., 2008). Second, in 4T1

mouse mammary carcinoma cells, TGF leads to SMAD1 andSMAD5 phosphorylation without the requirement of a BMP-like typeI receptor because the TRI receptor kinase can directly phosphorylateSMAD1/5 (Liu et al., 2008). These findings suggest that a re-evaluation of TGF family signaling is needed through the elucidationof the type I receptors and Smad pathways that function in specificphysiological and developmental contexts.

TGF family signaling via non-Smad signalingproteinsOther proteins mediate TGF signaling in addition to Smads (seealso Moustakas and Heldin, 2005), and here we describe themechanistically best-established examples (Fig. 1C). TRIIphosphorylates the polarity protein PAR6, which regulates the localdegradation of the RHOA small GTPase that controls the assemblyof intercellular tight junctions in mammalian cells (Ozdamar et al.,2005). As tight junctions disassemble, epithelial architecturedisintegrates, followed by de-differentiation known as the epithelial-to-mesenchymal transition (EMT), an important developmental anddisease-associated process that is regulated by TGF signaling(reviewed by Moustakas and Heldin, 2007). During EMT, in

S*

(S*)

T*

(T*)

(T*)

(T*)

(T*)

T*

(T*)

T*

S*

(S*)

S*S*

S*

SMAD2

SMAD3

SMAD1

SMAD5

SMAD8

e3

β-hNLS

SMAD4 β-hNLS NES

L3

SAD

SMAD6

SMAD7NLS ?

NES ? L3

β-hNLS

β-hNLS NES

L3

β-hNLS

β-hNLS

S*MS*

S*VS*

S*VS*

S*VS*

S*VS*

NES

UbSumo

Ac

S* S*

S* S*

(S*)

(S*)

(S*)

(S*)

R-Smads

Co-Smad

I-Smads NLS ?

NES ? L3

Ac

Ac

PY

PY

PY

PY

PY

PY

PY

Me

(S*)

S*S*T*

(S* S*T*S*)

T*

(S* S*T*S*)

A

B

C

L3

L3

L3

L3

Fig. 2. The Smad family. Simplified structures of the eight human Smad proteins divided into (A) Receptor-activated (R) Smads; (B) common-mediator (Co) Smad; and (C) inhibitory (I) Smads. The conserved N-terminal Mad-homology 1 (MH1) (blue) and C-terminal MH2 (green) domainsare shown. Highlighted are the nuclear localization signal (NLS, striped box); the two unique inserts of SMAD2 (triangles), the second of whichcorresponds to exon 3 (e3); the -hairpin domain that binds to DNA (-h, black box); the proline-tyrosine (PY) motif in the linker domain that isrecognized by the WW domain of Smurf family proteins; the Smad activation domain (SAD) at the linker-MH2 border; the nuclear export signal(NES, hatched box); and the L3 loop of the MH2 domain. Phosphorylatable serine and threonine residues are shown; S/T* indicates experimentallyproven phosphorylation sites; (S/T*) indicates a conserved residue with a predicted phosphorylation motif that awaits experimental validation. TheC-terminal serines that are phosphorylated by the type I receptor kinases are shown in green (S*VS*, S*MS*); red S/T residues are phosphorylatedby the MAPKs ERK1/2; brown S/T residues are phosphorylated by protein kinase C and by calmodulin-dependent kinase II; blue S/T residues arephosphorylated by cyclin-dependent kinases; and black S/T residues are phosphorylated by glycogen synthase kinase 3. Sumoylation (Sumo),ubiquitylation (Ub), methylation (Me) and acetylation (Ac) sites are indicated with colored arrowheads.

DEVELO

PMENT

Page 6: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3704

addition to tight junctions, adherens junctions and desmosomes ofpolarized epithelial cells are destroyed and remodeled to give rise tomesenchymal-like cells that are motile and invasive. It should benoted that, in addition to the above direct mechanism of tightjunction disassembly, TGF elicits EMT via Smad signaling,leading to the transcriptional induction of major inducers of thisdifferentiation process (Thuault et al., 2008).

Whereas the TGF-PAR6 pathway locally degrades RHOA in abreast epithelial cell culture model, other studies have demonstratedthe positive activation of Rho GTPase signaling by TGF and BMPreceptors in diverse cell types (reviewed by Kardassis et al., 2009).However, the mechanism of Rho activation by TGF receptorsremains unclear (Fig. 1C).

In a distinct mechanism, the TGF type I receptor phosphorylatesboth serine and tyrosine residues in the SHCA (SHC1) adaptor,which then recruits the adaptor protein GRB2 and the Ras guanineexchange factor (GEF) son of sevenless (SOS) in mammalian cells(Fig. 1C) (Lee, M. K. et al., 2007). This leads to activation of theRas-Raf-MEK-Erk mitogen-activated protein kinase (MAPK)signaling cascade, which can regulate cell proliferation or migration.Future work might decipher to what extent a specific biologicalresponse to TGF receptor signaling depends on its strongserine/threonine, or on its weaker tyrosine kinase, activity.

The tyrosine kinase Src can phosphorylate Tyr284 in thecytoplasmic domain of the TRII receptor, leading to GRB2 andSHC recruitment and to the activation of the p38 MAPK pathway(Galliher and Schiemann, 2007). Src-dependent TRIIphosphorylation regulates breast cancer cell proliferation andinvasiveness, possibly without affecting the Smad signaling output(Galliher-Beckley and Schiemann, 2008).

As a final example, TGF-induced receptor heterotetramersrecruit the ubiquitin ligase tumor necrosis factor receptor-associated factor 6 (TRAF6) to the TRI cytoplasmic domain inmammalian cells (Fig. 1C). TRAF6 ubiquitylates and activates thecatalytic activity of the TGF-activated kinase 1 (TAK1; MAP3K7),leading to activation of the p38 and c-Jun N-terminal kinase (JNK)cascades, which regulate apoptosis or cell migration (Sorrentino etal., 2008; Yamashita et al., 2008). The TRI kinase activity isdispensable for this pathway (Sorrentino et al., 2008).

The developmental significance of these non-Smad pathwaysremains to be elucidated. However, it has recently been shown thatboth p38 MAPK and Smad signaling play important rolesdownstream of TGF during mouse palate and tooth development(Xu, X. et al., 2008). In Xenopus, the adaptor protein TRAF4 haspositive signaling roles in mediating both the BMP and nodal signalsthat regulate neural crest differentiation and migration (Kalkan et al.,2009). TRAF4 is a substrate of the ubiquitin ligase Smadubiquitylation regulatory factor 1 (SMURF1), which polyubiquity-lates and promotes TRAF4 degradation, thus limiting the activity ofthe BMP and nodal pathways in the Xenopus neural plate. Signalingvia multiple effectors enables the TGF pathways to be controlled byother pathways, as we discuss below, through the mechanisms thatcontrol Smad function in different cell compartments.

TGF receptor regulation and endocytosisReceptor phosphorylation is important for TGF family signaltransduction, and thus receptor dephosphorylation might also beimportant. New evidence shows that TGF/nodal receptors arereciprocally regulated by B and B, two isoforms of regulatorysubunit B of the protein phosphatase 2A (PP2A). PP2A that containsthe B subunit positively, whereas PP2A that contains the Bsubunit negatively, regulates receptor signaling in Xenopus embryos

and in mammalian cells (Batut et al., 2008). The direct moleculartargets of PP2A and the serine or threonine residues that theydephosphorylate await further analysis.

In addition, TRI can be sumoylated by an as yet unknown sumo-ligase in mammalian cells (Fig. 3A), which enhances TGFsignaling by facilitating SMAD3 recruitment to the receptor forphosphorylation (Kang et al., 2008). TRI sumoylation mightprovide a docking site for an adaptor that mediates SMAD3 bindingto the receptor, might induce a conformational change in the receptorthat is required for SMAD3 binding, or might be coupled to theinternalization mechanism.

The activated TGF receptors are internalized via clathrin-coatedpits into early endosomes, where the receptors encounter Smadanchor for receptor activation (SARA; ZFYVE9), which facilitatesthe recruitment of SMAD2 and SMAD3 to TRI and theirsubsequent phosphorylation (Fig. 3A,B) (Tsukazaki et al., 1998).SARA binds SMAD2 and SMAD3, but not BMP R-Smads, and alsocontacts the TRI receptor. A homolog of SARA, endofin(ZFYVE16), plays a similar role in BMP pathways (Fig. 3D,E) (Shiet al., 2007). However, endofin can also operate in TGF pathwaysby scaffolding SMAD4 and by mediating the formation ofcomplexes of SMAD2, SMAD3 and SMAD4 in association with thetype I receptor (Chen et al., 2007).

In mammals, SARA cooperates with the cytoplasmicpromyelocytic leukemia (cPML) tumor suppressor protein, whichstabilizes the SARA-Smad complex (Fig. 3B) (Lin et al., 2004). Thisprocess also involves another adaptor, the PML competitor for TGIFassociation (PCTA), which binds to the nuclear homeodomainrepressor protein TGIF (5�TG3�-interacting factor) and retainscPML in the nucleus (Faresse et al., 2008). In response to TGF,PCTA releases cPML to translocate to the cytoplasm, reach SARAand promote TGF signaling.

The Drosophila ortholog, dSARA, plays a similar role tomammalian SARA during Dpp/Mad signaling (Bökel et al., 2006).Since dSARA has no other homolog in Drosophila, it might mediateboth Mad and dSmad2 signaling, although this remains to beestablished. In the epithelial cells of the Drosophila wing, whichundergo apical and basolateral differentiation, Dpp receptor-dSARAcomplexes reside in apically located endosomes and are preciselysegregated during cell division (Bökel et al., 2006). In this way, thedaughter cells receive equal numbers of signaling complexes, whichensures the conservation of signaling strength from mother todaughter cells.

Many additional cytoplasmic regulators of early TGF receptorsignaling have been described recently (Fig. 3B). Most notably,the small GTPase RAP2 inhibits activin/nodal receptor recycling,thus controlling receptor levels on the surface of Xenopusembryonic cells (Choi et al., 2008). During signaling, RAP2antagonizes the negative effects of SMAD7, thus positivelycontributing to nodal signal propagation and the onset ofgastrulation. In addition, RIN1, a RAB5 GEF, promotes TGFreceptor endocytosis and overall signaling, which contributes tothe pro-tumorigenic action of TGF in breast epithelial cells (Huet al., 2008). The PDZ-containing protein erbin (ERBB2IP) bindsto phosphorylated SMAD2, SMAD3, prevents their associationwith SMAD4 in mammalian cells, and produces opposite effectson signaling to SARA or endofin (Dai, F. et al., 2007). The proteinDapper 2 (DACT2) contributes to TGF receptordownregulation, which modulates nodal signaling in Xenopus,zebrafish, and mice (Su et al., 2007), although its partners andmechanism of action require further exploration. It would beinteresting to elucidate the mechanism that regulates the

REVIEW Development 136 (22)

DEVELO

PMENT

Page 7: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3705REVIEWDevelopment 136 (22)

recruitment of positive regulators, such as SARA, endofin andRIN1, and of negative regulators, such as RAP2, erbin andDapper 2, to the TGF receptor complex to ensure appropriate R-Smad phosphorylation and SMAD4 association.

Receptor endocytosis both controls the flow of signaling andregulates the availability of TGF ligand on the cell surface. An invitro kinetic analysis of TGF ligand bioavailability has shown thatconstitutive endocytosis of TRII depletes excess ligand (Clarke etal., 2009). This mechanism enables a cell to fine-tune the level ofsignaling growth factor on the cell surface.

The regulatory proteins described above highlight the linkbetween TGF signaling and the regulation of receptorinternalization. However, the developmental relevance of many ofthese factors awaits further analysis.

TGF receptor downregulation and the role of I-SmadsTGF ligand-receptor complexes can be additionally internalizedvia lipid rafts into caveolae, and then into lysosomes, where theligand-receptor complex is degraded (Di Guglielmo et al., 2003)(Fig. 3C). TGF receptor trafficking via caveolae is marked by theirassociation with I-Smads and SMURF1 or SMURF2 ubiquitinligases, which negatively regulate the signaling cascade.

The inhibitory Smads, SMAD6 and SMAD7, bind to type Ireceptors, thereby competitively inhibiting R-Smadphosphorylation and recruiting phosphatases and Smurf ubiquitinligases to downregulate receptor levels and function (reviewed byItoh and ten Dijke, 2007). Whereas SMAD7 inhibits both TGFand BMP pathways, SMAD6 more selectively inhibits BMPpathways (Fig. 1) and shows greater selectivity for the BMP typeI receptors ALK1, ALK2, ALK3 and ALK6, as demonstratedrecently in mammalian cells. Furthermore, SMAD6 binds witheven higher affinity to specific amino acid residues in theBMPRIA (ALK3) and BMPRIB (ALK6) kinase domains, than toALK1 and ALK2 domains (Goto et al., 2007). By contrast,SMAD7 shows broader specificity as it binds to all type Ireceptors via specific lysine residues in its MH2 domain(Mochizuki et al., 2004).

Two regulatory mechanisms that mediate the SMAD7-dependent ubiquitylation and downregulation of the TGFreceptor have recently been uncovered in mammalian cells (Fig.3C). The chaperone protein HSP90 binds to TRII and to TRIand protects them from ubiquitylation by SMURF2, positivelycontributing to TGF signaling (Wrighton et al., 2008).Conversely, the AMP-regulated kinase member salt-induciblekinase (SIK) is induced at the mRNA and protein levels by TGF

R-Smad

R-Smad

R-Smad

RI

P

RII

TGFβ

Cytoplasm

Nucleus

RI

P

RII

BMP

S Sumo-ligase

Receptor endocytosis

Endofin

A D

C

Receptor endocytosis

SaracPML

B

Receptor degradation

Cav1

CaveolaEndosome

E

Receptor degradation

Smad6Smurf

Caveola

F

Smad2P

Smad3P

Co-Smad

Smad5P

Smad1P

Co-Smad

Endosome

Smad7SmurfSIKHsp90

?

Rap2RinErbin

Smad8P

Chromatin

mRNA

TF

PP Co-Smad

mRNA

TF

PP Co-Smad

Chromatin

R-Smad

Fig. 3. TGF receptor endocytosis and downregulation. (A,D)The (A) TGF and (D) BMP receptor complexes are shown at the plasmamembrane. Type II receptors are labeled as RII, type I receptors as RI. Sumoylation (S) of the TRI that positively influences Smad signaling ishighlighted. (B)TGF receptors are internalized in endosomes characterized by the presence of the endocytic protein SARA and the regulatoryadaptor cPML. The receptors can signal by SMAD2 and SMAD3 phosphorylation and activation of the Co-Smad and by accumulation of nuclearSmad complexes on chromatin. Key regulatory proteins involved in the process of receptor endocytosis are listed on the side of the arrow thatindicates the flow of endosomal trafficking. (C)The TGF receptor degradation pathway via caveolae, which are characterized by the presence ofcaveolin 1 (CAV1), and the key regulatory proteins involved as shown within the box. (E)The corresponding endocytic pathway for BMP receptorswith the endocytic protein endofin is less well understood. (F)The corresponding caveolae-based degradation pathway for the BMP receptorsremains unexplored (?). Key established regulatory proteins are shown in the box. BMP (bone morphogenetic protein), cPML (cytoplasmicpromyelocytic leukemia protein), HSP90 (heat-shock protein of 90 kDa), RAP2 (Ras-related protein 2), RIN (Ras-like protein expressed in neurons),SARA (Smad anchor for receptor activation), SIK (salt-inducible kinase), Smurf (Smad ubiquitylation regulatory factor), TGF (transforming growthfactor ).

DEVELO

PMENT

Page 8: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3706

signaling, concomitantly with the induction of SMAD7 andSMURF2 (Kowanetz et al., 2008). SIK binds to SMAD7 and toTRI to promote receptor downregulation. The C. elegans SIKortholog, KIN-29, exhibits a conserved function by regulatingbody size in the Sma/Mab pathway; however, the molecularmechanism of KIN-29 action in worms remains unexplored(Maduzia et al., 2005).

Although SMAD7 primarily acts at the type I receptor level, italso resides in the nucleus, and new evidence suggests that it canbind to DNA and to nuclear complexes of SMAD2, SMAD3 andSMAD4, disrupting their complexes and inhibiting theirtranscriptional activity (Fig. 4) (Zhang et al., 2007).

Based on the importance of the mechanisms of TGF receptorendocytosis and downregulation, and the links of such mechanismsto the nucleocytoplasmic shuttling of Smads (see below), futurestudies into the biology of I-Smads promise to reveal interesting andnovel findings.

Regulation of Smad trafficking by motor proteinsIn parallel to TGF receptor endocytosis, R-Smads becomephosphorylated by the type I receptors and accumulate in thenucleus. However, Smads, like the receptors, show dynamicmobility and shuttle in and out of the nucleus even when they arenot activated by receptors (reviewed by Moustakas and Heldin,2008). The cytoplasmic trafficking of both TGF receptors andSmads is often mediated by motor proteins that are associatedwith microtubules. Motor proteins are important both before andafter R-Smad C-terminal phosphorylation.

Accordingly, Smads interact with kinesin 1, which mediates therecruitment of SMAD2 to the receptor complex in Xenopus andmammalian cells (Batut et al., 2007). Smads are sequestered awayfrom the receptors when bound to microtubules, from where theycan be released, for example, by connexin 43 (GJ1), whichcompetes with microtubules to bind Smads (Dai, P. et al., 2007).Additional motor proteins, such as the dynein light chain km23-1

(DYNLRB1), also promote Smad traffic towards the nucleus ofmammalian cells (Jin et al., 2007). The developmental roles of theconnexin 43 and km23-1 mechanisms have not yet beenspecifically addressed.

Microtubules also transport specialized pools of Smad proteins.For example, the pool of SMAD1 that has been subjected toinhibitory phosphorylation in its linker domain by Erk MAPKs(see below) moves towards the centrosome via microtubules,where it is degraded by proteasomes (Fuentealba et al., 2008).Interestingly, when mammalian embryonic and adult cells, orDrosophila blastoderm cells, complete mitosis, phosphorylatedSMAD1 and the centrosomal degrading apparatus segregate toonly one of the two daughter cells (Fuentealba et al., 2008). Thus,Smad trafficking and segregation to daughter cells is regulateddevelopmentally in a stage-specific manner. Additional studies inXenopus embryos show that SMAD2-SMAD4 complexes arerecruited to chromatin during every cell division, when mitosisdissolves the nuclear envelope (Saka et al., 2007). Again, thisevent is regulated developmentally as it does not occur prior to themid-blastula transition. Moreover, SMAD3 regulates the activityof the anaphase-promoting complex, a primary initiator of mitosisduring the mammalian cell cycle (Fujita et al., 2008). Whetherthese three mechanisms of Smad regulation during embryonic cellmitosis constitute one and the same process remains to beelucidated.

Mechanisms of Smad shuttling through thenuclear envelopeSmad nuclear import via nuclear pores is mediated bynucleoporins, which are integral constituents of the pores, and byimportins, carrier proteins that bind to both cargo proteins andnucleoporins and that catalyze their nuclear translocation in anenergy-dependent manner (reviewed by Moustakas and Heldin,2008). All Smads have a conserved NLS in their MH1 domain(Fig. 2), which binds to specific importins, such as importin (in

REVIEW Development 136 (22)

R-Smad

TF

R-SmadCo-Smad S

UP

U

P PU

SA

RI

P

RII

TGFβ

Co-Smad

Smad3P

Smad2P

Cytoplasm

Nucleus

p53HHMEts1TFAP2AIKKαBRG1

Smad7

mRNA

TF

Co-Smad

ArkadiaSmurf2

APC/Cdh1SnoN

Ski

mRNA

Chromatin

SnoNSki

Proteasomaldegradation

Smad3P

Smad2P

Fig. 4. Regulation of the nuclear Smadcomplex. Following TGF receptor activation, theactivated Smad complex translocates to thenucleus and associates with various Smad partners[represented by the blue, triangular transcriptionfactor (TF) and by examples discussed in the text]and undergoes post-translational modifications (A,acetylation; U, ubiquitylation; S, sumoylation; P,phosphorylation). Black arrows indicate a positiveeffect of a post-translational modification on Smadtranscriptional function, red T-bars indicate anegative effect. (Left) Derepression of Smad targetgenes when the SNON/SKI co-repressors areproteasomally degraded after being ubiquitylatedby the ubiquitin ligases arkadia, SMURF2 orAPC/CDH1. APC/CDH1 (anaphase-promotingcomplex and its ubiquitin ligase subunit CDH1),BRG1 (Brahma-related gene 1), ETS1 (v-etserythroblastosis virus E26 oncogene homolog 1),HHM (human homolog of Maid), IKK (IB kinase), SMURF2 (Smad ubiquitylation regulatory factor2), TFAP2A (transcription factor activatingenhancer-binding protein 2), TGF (transforminggrowth factor ).

DEVELO

PMENT

Page 9: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3707REVIEWDevelopment 136 (22)

the case of SMAD1 and SMAD3) and importin (in the case ofSMAD4). A recent genome-wide study in Drosophila S2 cellsreported the roles of additional importins, such as Msk, whichimports Mad, and of its mammalian orthologs, importin 7 andimportin 8, which import SMAD2, SMAD3 and SMAD4 (Xu etal., 2007; Yao et al., 2008).

Smad proteins have characterized nuclear export signals(NESs) in their MH2 (SMAD1, SMAD3) or linker (SMAD4)domains (Fig. 2). The NESs bind specific exportins – exportin 4for SMAD3 and exportin 1 for SMAD1 and SMAD4 – and exportis catalyzed by the small GTPase RAN. Recently, a new exportinwas described for SMAD2 and SMAD3, the RAN-bindingprotein 3 (RANBP3), which is a known exportin family member(Dai et al., 2009). RANBP3 was shown to recognizedephosphorylated nuclear SMAD2 and SMAD3 proteins andexport them to the cytoplasm. Importins and exportins forSMAD5 and SMAD8 or for the I-Smads remain to becharacterized. However, as we discuss below, our understandingof how Smads are transported into the nucleus has beensignificantly advanced by recent findings.

Regulation of Smad nuclear shuttlingThe dynamic movement of Smads in and out of the nucleus is highlyregulated. A recent mathematical model of SMAD2 and SMAD4trafficking reported that their nuclear accumulation in response toTGF reflects a shift in the equilibrium between the cytoplasmic andnuclear pools of Smads that is brought about by a decrease in nuclearexport (Schmierer et al., 2008). During signaling, however, low-level R-Smad dephosphorylation by nuclear phosphatases, amongother factors, continues to ensure their subsequent nuclear export(Schmierer et al., 2008).

R-Smad shuttling is regulated not only by cycles of receptor-mediated C-terminal phosphorylation and dephosphorylation bynuclear phosphatases (reviewed by Wrighton et al., 2009), butalso by sumoylation and ubiquitylation. Sumoylation of SMAD3by the protein inhibitor of activated Stat y (PIASy; PIAS4) sumo-ligase promotes its nuclear export in mammalian cells (Imoto etal., 2008). Sumoylation of Medea, by an as yet unidentified sumo-ligase, also promotes its nuclear export, providing negativeregulation that restricts the competence of early Drosophilaembryonic cells to respond to Dpp (Miles et al., 2008). Thismechanism resembles the previously established role ofmammalian SMAD4 sumoylation by PIAS ligases (reviewed byLönn et al., 2009). It is possible that in vivo Smad sumoylationmight not negatively regulate TGF signaling; rather, it mightpromote continuous Smad shuttling. However, underoverexpression conditions, sumo-ligases may shift the shuttlingequilibrium by pushing Smads to the cytoplasm, thus reducingtheir time in the nucleus.

SMAD4 can also be monoubiquitylated (Morén et al., 2003) bythe nuclear ubiquitin ligase TIF1 (ectodermin; TRIM33), whichpromotes its nuclear export and inhibits the formation of nuclearcomplexes of SMAD2, SMAD3 and SMAD4 (Dupont et al., 2009).Once monoubiquitylated, SMAD4 is exported from the nucleus(Wang et al., 2008), whereupon fat facets in mouse (FAM; USP9X)deubiquitylates it, recharging it for subsequent cycles of shuttling,as demonstrated in Drosophila, Xenopus and human cells (Dupontet al., 2009).

Other mechanisms also regulate the nuclear residence andfunction of Smad complexes. Heteromeric complexes of SMAD2,SMAD3 and SMAD4 bind to the shuttling protein transcriptionalco-activator with PDZ-binding motif (TAZ) in the nucleus of

human cells, and are then recruited to chromatin via factors suchas the activator-recruited co-factor (ARC) protein ARC105(MED15), a member of the Mediator complex that ensures theprogression of gene transcription (Varelas et al., 2008). TAZ isregulated by phosphorylation and by interaction with 14-3-3family adaptors that control its timely residence in the nucleus.Furthermore, the Drosophila nuclear lamin Otefin interacts withMedea and tethers Smad complexes to the nuclear envelope(Jiang et al., 2008). The Otefin-Medea complexes bind to specificgene-regulatory elements that control germline stem celldevelopment (Jiang et al., 2008). Future research into Smadshuttling and the regulation of Smad compartmentalization willbring to light additional regulatory mechanisms of TGFsignaling.

Negative regulation of Smad signaling byphosphorylation and ubiquitylationIn addition to regulating Smad nucleocytoplasmic shuttling, C-terminal tail dephosphorylation, linker domain phosphorylation andubiquitylation are implicated in the negative regulation of Smadsignaling (Figs 2 and 4).

The developmental importance of such post-translationalmodifications has been recognized during BMP-dependentneurogenesis in Xenopus and mouse C2C12 osteoblastdifferentiation. Fibroblast growth factor (FGF) signaling via Ras-Erk MAPK negatively regulates BMP signaling, as Erk (and GSK3kinase) directly phosphorylates the SMAD1 linker, leading torecruitment of SMURF1 and to the proteasomal degradation ofSMAD1 in perinuclear centrosomes (Fuentealba et al., 2007;Sapkota et al., 2007). During Xenopus neurogenesis, SMAD1degradation is triggered in three ways: by chordin antagonisingBMP activity extracellularly; by the Wnt antagonist Dickkopf 1blocking Wnt activity extracellularly; and by Erk MAPK pathwayactivation via FGF and insulin-like growth factor (IGF) signaling.Thus, FGF/IGF signaling provides negative feedback to BMPsignaling in the developing nervous system. Conversely, Wntsignaling induces GSK3 degradation, leading to decreasedSMAD1 linker phosphorylation and to its prolonged signaling, anevent that is required for Xenopus epidermal differentiation(Fuentealba et al., 2007). This is a good example of positive cross-talk between Wnt and BMP signaling during skin differentiation.

GSK3 also negatively controls TGF signaling, as it directlyphosphorylates SMAD3 in its MH1 domain in mammalian cells(Guo et al., 2008a). This phosphorylation is followed by theubiquitylation and proteasomal degradation of SMAD3, whichregulate its steady-state levels. By contrast, upon TGF receptorphosphorylation, SMAD3 can be further phosphorylated in its MH2domain by casein kinase 1 2, which leads to the specificubiquitylation and degradation of the activated form of SMAD3(Guo et al., 2008b).

A recent report has shed more light on the complexity of Smadregulation through the phosphorylation of its linker domain (Wanget al., 2009). After SMAD3 C-terminal phosphorylation by TRI inmammalian cells, nuclear GSK3 and cyclin-dependent kinasesphosphorylate three distinct SMAD3 linker residues,downregulating its transcriptional activity. Thus, TGF signalingtightly controls the activity of one of its main transducers throughhighly regulated phosphorylation events.

Recent evidence also shows that during mitosis of mammaliancells in culture, the kinase MPS1 (TTK) can directly C-terminallyphosphorylate SMAD2 and SMAD3, thus activating their nuclearactivities in the absence of TGF receptor activation (Zhu et al., D

EVELO

PMENT

Page 10: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3708

2007). This is one of the first clear TGF-independent mechanismsthat engages Smads and mimics the action of TGF. Thus, whereasR-Smad C-terminal phosphorylation by type I receptor kinases is apositive regulator of TGF family signaling, Smad phosphorylationby other kinases can negatively affect TGF family signaling in acell cycle-, developmental- or tissue-specific manner.

Transcriptional regulation by SmadsThe list of transcription factors to which Smads bind to regulate geneexpression continues to grow (see Table S1 in the supplementarymaterial) (reviewed by Feng and Derynck, 2005). Nuclear Smadcomplexes bind with weak affinity to Smad-binding elements(SBEs) on DNA (reviewed by Schmierer and Hill, 2007). Notably,the most common isoform of SMAD2 fails to bind to SBEs owingto an insertion within its DNA-binding domain, which resides in theMH1 domain of all Smads (see Fig. 2). SMAD3 recognizes 5�-GTCTG-3� as its SBE. By contrast, the BMP Smads and SMAD4recognize GC-rich sequences that have less conserved motifs, whichare sometimes in close proximity to an SBE. In general, recruitmentof Smad complexes to chromatin is dependent on their directinteraction with transcription factors that bind to DNA with higheraffinity (Fig. 4).

Upon binding to DNA and to their transcriptional partners, Smadsrecruit co-activators and histone acetyltransferases, such as p300,C/EBP-binding protein (CBP) and p300/CBP-associated factor(P/CAF), facilitating the initiation of transcription (reviewed bySchmierer and Hill, 2007). Recent evidence has shown thatp300/CBP also acetylates SMAD2/3, enhancing their DNA-bindingactivity in mammalian cells (Simonsson et al., 2006; Tu and Luo,2007). Conversely, histone deacetylases inhibit SMAD1transcriptional activity during neuronal differentiation in the mouseembryonic brain (Shakéd et al., 2008). However, direct acetylationor deacetylation of BMP-specific Smads has yet to be demonstrated.

The negative regulation of Smad signaling by Smadubiquitylation was summarized above. Positive regulation of nuclearSmad signaling by ubiquitylation has more recently come to lightfrom studies in mouse embryos and in mammalian cells (Mavrakiset al., 2007). Nuclear Smad complexes associate with the ubiquitinligase arkadia (RNF111) in a ligand-dependent manner and promotethe ubiquitylation and degradation of their interacting co-repressorsSKI and SNON (SKIL) (Le Scolan et al., 2008; Levy et al., 2007;Nagano et al., 2007). This mechanism brings about the derepressionand transcriptional induction of target genes by nuclear Smads (Fig.4). Interestingly, the proteasomal degradation of SNON depends onits phosphorylation by TAK1, the non-Smad effector of TGFsignaling (see Fig. 1C) (Kajino et al., 2007). However, it is unclearwhether SNON ubiquitylation by arkadia requires its priorphosphorylation by TAK1. Arkadia also ubiquitylates the inhibitorySMAD7 (Koinuma et al., 2003), but whether this process takes placein the nucleus or in the cytoplasm awaits clarification.

Genome-wide screens have revealed an association betweenSmads and the SWI/SNF family chromatin remodeling proteinBrahma-related gene 1 (BRG1; SMARCA4) and the DNA-bindingproteins ETS1 and transcription factor activating enhancer-bindingprotein 2 (TFAP2) (Koinuma et al., 2009; Xi et al., 2008). Acurrent model suggests that chromatin-bound Smads cannot performtranscriptional work in the absence of essential chromatinremodeling factors, such as BRG1 and the mediator componentARC105 (reviewed by Schmierer and Hill, 2007). Interestingly,ARC105 localization in distinct chromatin domains is regulated byTAZ, the nuclear Smad-tethering factor (Varelas et al., 2008). Theseearly reports open the door to future studies that might demonstrate

how TGF alters the dynamic architecture of chromatin, leading togene-specific transcriptional induction or repression. Such researchmight, for the first time, establish links between the epigeneticregulation of chromatin and the function of the TGF pathways.

Regulatory mechanisms of Smad transcriptionalco-factorsFrom the numerous Smad-transcription factor complexes and theirresulting mechanisms of target gene regulation (see Table S1 in thesupplementary material) (Feng and Derynck, 2005), we highlighthere a few selected examples that are of demonstrated or potentialdevelopmental relevance.

Xenopus mesoderm specification is driven by the concerted actionof TGF/activin and FGF-Ras-Erk MAPK signaling (Cordenonsi etal., 2007). The FGF-Ras-Erk MAPK pathway acts in distinct regionsof the developing Xenopus embryo, such as in the marginal zone,and induces, via phosphorylation, the activity of casein kinases,which then phosphorylate serines 6 and 9 of the tumor suppressorp53, contributing to mesoderm development. This phosphorylationactivates the transcriptional activity of p53, making it competent topair with Smads. This interaction leads to the transcriptionalinduction of mesoderm-defining genes, such as the transcriptionfactors Snail, Xbra (brachyury) and Mix.2. Conversely, duringXenopus ectoderm specification, p53 is inhibited by the zinc-fingerprotein XFDL156 (Sasai et al., 2008). This mechanism is essentialfor preventing the aberrant activation of nodal signaling in theectoderm, the developmental fate of which depends primarily on theactivity of BMP pathways. Although the above example emphasizespositive cross-talk between FGF and activin signaling duringXenopus mesoderm specification, this should not be interpreted asthe only developmental FGF signaling mechanism during frogmesoderm induction. The FGF response is multifactorial andmultigenic, as revealed by recent genome-wide transcriptomicscreens (Branney et al., 2009). Interestingly, although Smadscooperate with wild-type p53 to promote developmental processes,they also cooperate with mutant p53, which often accumulates inhuman cancers (Adorno et al., 2009; Kalo et al., 2007). The Smad-mutant p53 complex represses TGFBR2 transcription, leading to theinduction of pro-metastatic genes.

An unexpected transcriptional partner of SMAD2 and SMAD3 isthe well-characterized IB kinase (IKK; CHUK), whichparticipates in the nuclear factor B (NFB) pathway (Descargueset al., 2008). In the epidermis of Smad4-null mice, a complex ofSMAD2, SMAD3 and IKK forms in the absence of SMAD4 andregulates mammalian keratinocyte differentiation by binding to theregulatory sequences of the transcription factor genes Mad1 (Mxd1– Mouse Genome Informatics) and Ovol1, which induce epidermaldifferentiation. Transcriptional induction mediated by the complexof SMAD2, SMAD3 and IKK is independent of the kinase activityof IKK. Interestingly, invasive squamous cell carcinomas that areresistant to the tumor suppressor action of TGF have defectiveIKK that cannot enter the nucleus and act as a Smad co-factor(Marinari et al., 2008).

Genes that are either transcriptionally co-regulated at the samedevelopmental time, or in the same tissue, are referred to assynexpression groups (Niehrs and Pollet, 1999). One such group isthe inhibitor of differentiation (Id) family of genes, which like otherTGF-responsive synexpression groups, respond to TGF familymembers via specific regulatory sequences present in their genes(see Table S1 in the supplementary material) (Karaulanov et al.,2004). Synexpression groups also require specific Smad-interactingtranscriptional co-factors for their expression. For example, the

REVIEW Development 136 (22)

DEVELO

PMENT

Page 11: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3709REVIEWDevelopment 136 (22)

mammalian FoxO transcription factors bind to Smads andcoordinate the regulation of 11 genes that define the cytostatic,apoptotic and adaptive signaling responses of keratinocytes to TGF(see Table S1 in the supplementary material) (Gomis et al., 2006a).The helix-loop-helix HHM (human homolog of Maid; CCNDBP1)protein regulates a specific synexpression group of cell cycle andcell migration regulators (see Table S1 in the supplementarymaterial), and, accordingly, regulates growth inhibition andmigration in mammalian epithelial cells in response to TGF, whilemediating other responses in different cells (Ikushima et al., 2008).

In the Drosophila wing imaginal disc, the BMP ligand Dppactivates Mad (an R-Smad) and Medea (its Co-Smad) (see Table 1),which then directly repress certain transcription factor genes,including the transcriptional repressor brinker (brk). This repressionof brk by Mad-Medea leads to the derepression of optomotor blind(omb; bifid – FlyBase), zerknüllt (zen) and spalt (sal), which encodetranscription factors that regulate the expression of othertranscription factors, to provide patterning and morphogeneticinformation to the developing wing (de Celis and Barrio, 2000; Shenet al., 2008). sal, however, additionally requires direct binding andtransactivation by the Mad-Medea complex.

In C. elegans, the sma-9 gene is involved in neuronalspecification within a restricted group of rays in the tail of thedeveloping worm and is expressed during early larval stages (Lianget al., 2003). SMA-9 is the ortholog of the Drosophila zinc-fingertranscription factor Schnurri, a co-factor of the Mad-Medea complexthat represses brk expression (Marty et al., 2000). By analogy withDrosophila, SMA-9 might mediate BMP-like DBL-1 signaling bycomplexing with SMA-2, SMA-3 or SMA-4. Indeed, SMA-9 actsas both a transcriptional repressor and an activator downstream ofDBL-1. Newly identified targets of this pathway are orthologs oftranscription factors that are already implicated in mammalian BMPsignaling, such as Runx and Fos, or orthologs of Hedgehog signalingproteins (Liang et al., 2007). Interestingly, mouse Schnurri-2(HIVEP2 – Mouse Genome Informatics) also regulatesTGF/BMP-dependent gene expression. It binds to SMAD1-SMAD4 and to the transcriptional co-factor C/EBP(CCAAT/enhancer-binding protein ) to induce the PPAR2(peroxisome proliferator-activated receptor 2; Pparg) gene thatregulates adipocyte differentiation (Jin et al., 2006). Thus, mice thatlack the Schnurri-2 gene have reduced fat.

Finally, as we discuss in Box 3, TGF family transcriptionalregulation in development also occurs via the regulation of micro-RNA (miRNA) genes, and via the reciprocal regulation of TGFsignaling by miRNAs.

ConclusionsAs TGF research continues with ever increasing speed, we foreseeimportant novel findings regarding the mechanisms of TGFreceptor regulation and specificity of signaling, cytoplasmictrafficking of receptors and Smads, nuclear dynamics of Smad-chromatin associations and their relationship to developmentalprocesses. The functional implications of ‘promiscuous’ signalingby TGF family receptor kinases that simultaneously activateTGF- and BMP-like Smad pathways, as well as MAPK and otherpathways, needs to be analyzed carefully and with quantitativemethods. The area of post-translational modifications of TGFreceptors and Smads will continue its prolific expansion. Moresensitive proteomic approaches will be useful in dissecting all thecomponents of signaling complexes and their dynamic nature,especially if coupled to multi-protein imaging in real time. Progressin the modeling of signaling dynamics and of the protein networks

that participate in the TGF family cascades should provide freshideas about new regulatory nodes in the network, and should alsodefine more quantitatively critical parameters that govern thebehavior of the network. A major challenge is to decipher the rolesof the TGF pathways during late stages of embryogenesis andduring neonatal life by conditional activation and inactivation ofTGF signaling components in model organisms. The importanceof cross-talk during different developmental stages between TGFand Wnt, Hedgehog, FGF or other pathways should be another focusfor future research. Finally, in the context of development, morecomplete circuits of target genes, and their corresponding protein orRNA regulators, will need to be delineated in the global effort toprovide a systems-level description of TGF pathways in everytissue and organ.

AcknowledgementsWe acknowledge funding by the Ludwig Institute for Cancer Research (LICR),the Atlantic Philanthropies/LICR Clinical Discovery Program, the SwedishCancer Society, the Swedish Research Council, the European Union MarieCurie Research Training Network ‘EpiPlastCarcinoma’ and the European UnionNetwork of Excellence ‘ENFIN’. We acknowledge the contribution of all pastand present members of the TGF signaling group to the scientific workemanating from our laboratory.

Supplementary materialSupplementary material for this article is available athttp://dev.biologists.org/cgi/content/full/136/22/3699/DC1

ReferencesAdorno, M., Cordenonsi, M., Montagner, M., Dupont, S., Wong, C., Hann,

B., Solari, A., Bobisse, S., Rondina, M. B., Guzzardo, V. et al. (2009). AMutant-p53/Smad complex opposes p63 to empower TGF-induced metastasis.Cell 137, 87-98.

Affolter, M. and Basler, K. (2007). The Decapentaplegic morphogen gradient:from pattern formation to growth regulation. Nat. Rev. Genet. 8, 663-674.

Box 3. TGF/BMP signaling and microRNAsTGF regulates micro-RNA (miRNA) gene expression in mammaliancells (Zavadil et al., 2007). TGF promotes the epithelial-to-mesenchymal transition (EMT) by repressing transcription of the miR-200 family (Burk et al., 2008; Gregory et al., 2008; Korpal et al.,2008), which downregulates the pro-EMT transcription factors zinc-finger E-box-binding homeobox 1 (ZEB1) and ZEB2. Transcriptionalrepression of miR-24, a positive regulator of myogenesis, in partexplains the anti-myogenic effects of TGF (Sun et al., 2008). BMPinduces osteoblast differentiation from mesenchymal progenitorsand concomitantly regulates expression of 22 miRNAs (Li et al.,2008). Among these, miR-133 targets the transcription factorRUNX2, a known target of BMP/Smad signaling that promotesosteoblast differentiation, whereas miR-135 targets SMAD5. ThesemiRNAs progressively inhibit BMP signaling to fine-tune bonedevelopment.

In addition, SMAD1, SMAD3 and SMAD5, but not SMAD4,directly bind to and regulate components of the DROSHAmicroprocessor complex, which regulates miRNA biogenesis (Daviset al., 2008). This post-transcriptional mechanism selectivelyregulates only certain miRNAs during vascular smooth muscledifferentiation. Understanding the mechanism of this selectivity andits developmental relevance is important.

miRNAs also target TGF signaling components. In Xenopus, miR-15 and miR-16 downregulate ActRIIA, which restricts receptorexpression to specific cells, thus defining the embryonic territory thatresponds to activin/nodal (Martello et al., 2007). In zebrafish, miR-430 downregulates the nodal-like ligand squint (ndr1 – ZebrafishInformation Network) and its extracellular antagonist lefty, which isrequired for proper patterning of the early fish embryo (Choi et al.,2007).

DEVELO

PMENT

Page 12: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3710

Anttonen, M., Parviainen, H., Kyronlahti, A., Bielinska, M., Wilson, D. B.,Ritvos, O. and Heikinheimo, M. (2006). GATA-4 is a granulosa cell factoremployed in inhibin-alpha activation by the TGF- pathway. J. Mol. Endocrinol.36, 557-568.

Bangi, E. and Wharton, K. (2006). Dpp and Gbb exhibit different effective rangesin the establishment of the BMP activity gradient critical for Drosophila wingpatterning. Dev. Biol. 295, 178-193.

Barrio, R. and de Celis, J. F. (2004). Regulation of spalt expression in theDrosophila wing blade in response to the Decapentaplegic signaling pathway.Proc. Natl. Acad. Sci. USA 101, 6021-6026.

Batut, J., Howell, M. and Hill, C. S. (2007). Kinesin-mediated transport of Smad2is required for signaling in response to TGF- ligands. Dev. Cell 12, 261-274.

Batut, J., Schmierer, B., Cao, J., Raftery, L. A., Hill, C. S. and Howell, M.(2008). Two highly related regulatory subunits of PP2A exert opposite effects onTGF-/Activin/Nodal signalling. Development 135, 2927-2937.

Benchabane, H. and Wrana, J. L. (2003). GATA- and Smad1-dependentenhancers in the Smad7 gene differentially interpret bone morphogeneticprotein concentrations. Mol. Cell. Biol. 23, 6646-6661.

Bianco, C., Normanno, N., Salomon, D. S. and Ciardiello, F. (2004). Role of thecripto (EGF-CFC) family in embryogenesis and cancer. Growth Factors 22, 133-139.

Blanchet, M. H., Le Good, J. A., Mesnard, D., Oorschot, V., Baflast, S.,Minchiotti, G., Klumperman, J. and Constam, D. B. (2008). Cripto recruitsFurin and PACE4 and controls Nodal trafficking during proteolytic maturation.EMBO J. 27, 2580-2591.

Blokzijl, A., ten Dijke, P. and Ibanez, C. F. (2002). Physical and functionalinteraction between GATA-3 and Smad3 allows TGF-beta regulation of GATAtarget genes. Curr. Biol. 12, 35-45.

Blokzijl, A., Dahlqvist, C., Reissmann, E., Falk, A., Moliner, A., Lendahl, U.and Ibanez, C. F. (2003). Cross-talk between the Notch and TGF- signalingpathways mediated by interaction of the Notch intracellular domain withSmad3. J. Cell Biol. 163, 723-728.

Bökel, C., Schwabedissen, A., Entchev, E., Renaud, O. and González-Gaitán,M. (2006). Sara endosomes and the maintenance of Dpp signaling levels acrossmitosis. Science 314, 1135-1139.

Branney, P. A., Faas, L., Steane, S. E., Pownall, M. E. and Isaacs, H. V. (2009).Characterisation of the fibroblast growth factor dependent transcriptome inearly development. PLoS ONE 4, e4951.

Brodin, G., Ahgren, A., ten Dijke, P., Heldin, C.-H. and Heuchel, R. (2000).Efficient TGF- induction of the Smad7 gene requires cooperation between AP-1, Sp1, and Smad proteins on the mouse Smad7 promoter. J. Biol. Chem. 275,29023-29030.

Brown, C. O., 3rd, Chi, X., Garcia-Gras, E., Shirai, M., Feng, X. H. andSchwartz, R. J. (2004). The cardiac determination factor, Nkx2-5, is activated bymutual cofactors GATA-4 and Smad1/4 via a novel upstream enhancer. J. Biol.Chem. 279, 10659-10669.

Bruna, A., Darken, R. S., Rojo, F., Ocana, A., Penuelas, S., Arias, A., Paris, R.,Tortosa, A., Mora, J., Baselga, J. et al. (2007). High TGF-Smad activityconfers poor prognosis in glioma patients and promotes cell proliferationdepending on the methylation of the PDGF-B gene. Cancer Cell 11, 147-160.

Burk, U., Schubert, J., Wellner, U., Schmalhofer, O., Vincan, E., Spaderna, S.and Brabletz, T. (2008). A reciprocal repression between ZEB1 and members ofthe miR-200 family promotes EMT and invasion in cancer cells. EMBO Rep. 9,582-589.

Chen, C. and Shen, M. M. (2004). Two modes by which Lefty proteins inhibitnodal signaling. Curr. Biol. 14, 618-624.

Chen, C. R., Kang, Y., Siegel, P. M. and Massague, J. (2002). E2F4/5 and p107as Smad cofactors linking the TGFbeta receptor to c-myc repression. Cell 110,19-32.

Chen, D. and McKearin, D. M. (2003). A discrete transcriptional silencer in thebam gene determines asymmetric division of the Drosophila germline stem cell.Development 130, 1159-1170.

Chen, X., Rubock, M. J. and Whitman, M. (1996). A transcriptional partner forMAD proteins in TGF- signalling. Nature 383, 691-696.

Chen, Y., Blom, I. E., Sa, S., Goldschmeding, R., Abraham, D. J. and Leask, A.(2002). CTGF expression in mesangial cells: involvement of SMADs, MAP kinase,and PKC. Kidney Int. 62, 1149-1159.

Chen, Y. G., Wang, Z., Ma, J., Zhang, L. and Lu, Z. (2007). Endofin, a FYVEdomain protein, interacts with Smad4 and facilitates transforming growthfactor- signaling. J. Biol. Chem. 282, 9688-9695.

Chen, X., Xu, H., Yuan, P., Fang, F., Huss, M., Vega, V. B., Wong, E., Orlov, Y.L., Zhang, W., Jiang, J. et al. (2008). Integration of external signaling pathwayswith the core transcriptional network in embryonic stem cells. Cell 133, 1106-1117.

Chi, X. Z., Yang, J. O., Lee, K. Y., Ito, K., Sakakura, C., Li, Q. L., Kim, H. R.,Cha, E. J., Lee, Y. H., Kaneda, A. et al. (2005). RUNX3 suppresses gastricepithelial cell growth by inducing p21WAF1/Cip1 expression in cooperation withtransforming growth factor -activated SMAD. Mol. Cell. Biol. 25, 8097-8107.

Cho, G., Lim, Y., Zand, D. and Golden, J. A. (2008). Sizn1 is a novel protein that

functions as a transcriptional coactivator of bone morphogenic protein signaling.Mol. Cell. Biol. 28, 1565-1572.

Choi, S. C., Kim, G. H., Lee, S. J., Park, E., Yeo, C. Y. and Han, J. K. (2008).Regulation of activin/nodal signaling by Rap2-directed receptor trafficking. Dev.Cell 15, 49-61.

Choi, W. Y., Giraldez, A. J. and Schier, A. F. (2007). Target protectors revealdampening and balancing of Nodal agonist and antagonist by miR-430. Science318, 271-274.

Chou, W. C., Prokova, V., Shiraishi, K., Valcourt, U., Moustakas, A.,Hadzopoulou-Cladaras, M., Zannis, V. I. and Kardassis, D. (2003).Mechanism of a transcriptional cross talk between transforming growth factor--regulated Smad3 and Smad4 proteins and orphan nuclear receptorhepatocyte nuclear factor-4. Mol. Biol. Cell 14, 1279-1294.

Clarke, D. C., Brown, M. L., Erickson, R. A., Shi, Y. and Liu, X. (2009).Transforming growth factor depletion is the primary determinant of Smadsignaling kinetics. Mol. Cell. Biol. 29, 2443-2455.

Cohen, T. V., Kosti, O. and Stewart, C. L. (2007). The nuclear envelope proteinMAN1 regulates TGF signaling and vasculogenesis in the embryonic yolk sac.Development 134, 1385-1395.

Cordenonsi, M., Dupont, S., Maretto, S., Insinga, A., Imbriano, C. andPiccolo, S. (2003). Links between tumor suppressors: p53 is required for TGF-gene responses by cooperating with Smads. Cell 113, 301-314.

Cordenonsi, M., Montagner, M., Adorno, M., Zacchigna, L., Martello, G.,Mamidi, A., Soligo, S., Dupont, S. and Piccolo, S. (2007). Integration of TGF- and Ras/MAPK signaling through p53 phosphorylation. Science 315, 840-843.

Costamagna, E., Garcia, B. and Santisteban, P. (2004). The functionalinteraction between the paired domain transcription factor Pax8 and Smad3 isinvolved in transforming growth factor- repression of the sodium/iodidesymporter gene. J. Biol. Chem. 279, 3439-3446.

Dahlqvist, C., Blokzijl, A., Chapman, G., Falk, A., Dannaeus, K., Ibanez, C. F.and Lendahl, U. (2003). Functional Notch signaling is required for BMP4-induced inhibition of myogenic differentiation. Development 130, 6089-6099.

Dai, F., Chang, C., Lin, X., Dai, P., Mei, L. and Feng, X. H. (2007). Erbin inhibitstransforming growth factor signaling through a novel Smad-interactingdomain. Mol. Cell. Biol. 27, 6183-6194.

Dai, F., Lin, X., Chang, C. and Feng, X. H. (2009). Nuclear export of Smad2 andSmad3 by RanBP3 facilitates termination of TGF- signaling. Dev. Cell 16, 345-357.

Dai, P., Nakagami, T., Tanaka, H., Hitomi, T. and Takamatsu, T. (2007). Cx43mediates TGF- signaling through competitive Smads binding to microtubules.Mol. Biol. Cell 18, 2264-2273.

Daly, A. C., Randall, R. A. and Hill, C. S. (2008). Transforming growth factor -induced Smad1/5 phosphorylation in epithelial cells is mediated by novelreceptor complexes and is essential for anchorage-independent growth. Mol.Cell. Biol. 28, 6889-6902.

Datta, P. K., Blake, M. C. and Moses, H. L. (2000). Regulation of plasminogenactivator inhibitor-1 expression by transforming growth factor--induced physicaland functional interactions between smads and Sp1. J. Biol. Chem. 275, 40014-40019.

Davis, B. N., Hilyard, A. C., Lagna, G. and Hata, A. (2008). SMAD proteinscontrol DROSHA-mediated microRNA maturation. Nature 454, 56-61.

de Celis, J. F. and Barrio, R. (2000). Function of the spalt/spalt-related genecomplex in positioning the veins in the Drosophila wing. Mech. Dev. 91, 31-41.

De Robertis, E. M. and Kuroda, H. (2004). Dorsal-ventral patterning and neuralinduction in Xenopus embryos. Annu. Rev. Cell Dev. Biol. 20, 285-308.

Degnin, C., Jean, F., Thomas, G. and Christian, J. L. (2004). Cleavages withinthe prodomain direct intracellular trafficking and degradation of mature bonemorphogenetic protein-4. Mol. Biol. Cell 15, 5012-5020.

Derynck, R. and Miyazono, K. (2008). TGF- and the TGF- family. In The TGF-Family (ed. R. Derynck and K. Miyazono), pp. 29-43. Cold Spring Harbor, NY:Cold Spring Harbor Laboratory Press.

Descargues, P., Sil, A. K., Sano, Y., Korchynskyi, O., Han, G., Owens, P.,Wang, X. J. and Karin, M. (2008). IKK is a critical coregulator of a Smad4-independent TGF-Smad2/3 signaling pathway that controls keratinocytedifferentiation. Proc. Natl. Acad. Sci. USA 105, 2487-2492.

Di Guglielmo, G. M., Le Roy, C., Goodfellow, A. F. and Wrana, J. L. (2003).Distinct endocytic pathways regulate TGF- receptor signalling and turnover.Nat. Cell Biol. 5, 410-421.

Dick, A., Hild, M., Bauer, H., Imai, Y., Maifeld, H., Schier, A. F., Talbot, W. S.,Bouwmeester, T. and Hammerschmidt, M. (2000). Essential role of Bmp7(snailhouse) and its prodomain in dorsoventral patterning of the zebrafishembryo. Development 127, 343-354.

Dupont, S., Mamidi, A., Cordenonsi, M., Montagner, M., Zacchigna, L.,Adorno, M., Martello, G., Stinchfield, M. J., Soligo, S., Morsut, L. et al.(2009). FAM/USP9x, a deubiquitinating enzyme essential for TGF signaling,controls Smad4 monoubiquitination. Cell 136, 123-135.

Eimon, P. M. and Harland, R. M. (2002). Effects of heterodimerization andproteolytic processing on Derrière and Nodal activity: implications for mesoderminduction in Xenopus. Development 129, 3089-3103.

Faresse, N., Colland, F., Ferrand, N., Prunier, C., Bourgeade, M. F. and Atfi, A.

REVIEW Development 136 (22)

DEVELO

PMENT

Page 13: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3711REVIEWDevelopment 136 (22)

(2008). Identification of PCTA, a TGIF antagonist that promotes PML function inTGF- signalling. EMBO J. 27, 1804-1815.

Feng, X. H. and Derynck, R. (2005). Specificity and versatility in TGF- signalingthrough Smads. Annu. Rev. Cell Dev. Biol. 21, 659-693.

Feng, X. H., Lin, X. and Derynck, R. (2000). Smad2, Smad3 and Smad4cooperate with Sp1 to induce p15Ink4B transcription in response to TGF-. EMBOJ. 19, 5178-5193.

Finnson, K. W., Parker, W. L., ten Dijke, P., Thorikay, M. and Philip, A. (2008).ALK1 opposes ALK5/Smad3 signaling and expression of extracellular matrixcomponents in human chondrocytes. J. Bone Miner. Res. 23, 896-906.

Fuentealba, L. C., Eivers, E., Ikeda, A., Hurtado, C., Kuroda, H., Pera, E. M.and De Robertis, E. M. (2007). Integrating patterning signals: Wnt/GSK3regulates the duration of the BMP/Smad1 signal. Cell 131, 980-993.

Fuentealba, L. C., Eivers, E., Geissert, D., Taelman, V. and De Robertis, E. M.(2008). Asymmetric mitosis: unequal segregation of proteins destined fordegradation. Proc. Natl. Acad. Sci. USA 105, 7732-7737.

Fujita, T., Epperly, M. W., Zou, H., Greenberger, J. S. and Wan, Y. (2008).Regulation of the anaphase-promoting complex-separase cascade bytransforming growth factor- modulates mitotic progression in bone marrowstromal cells. Mol. Biol. Cell 19, 5446-5455.

Furumatsu, T., Tsuda, M., Taniguchi, N., Tajima, Y. and Asahara, H. (2005).Smad3 induces chondrogenesis through the activation of SOX9 via CREB-binding protein/p300 recruitment. J. Biol. Chem. 280, 8343-8350.

Galliher, A. J. and Schiemann, W. P. (2007). Src phosphorylates Tyr284 in TGF-type II receptor and regulates TGF- stimulation of p38 MAPK during breastcancer cell proliferation and invasion. Cancer Res. 67, 3752-3758.

Galliher-Beckley, A. J. and Schiemann, W. P. (2008). Grb2 binding to Tyr284 inTR-II is essential for mammary tumor growth and metastasis stimulated by TGF-. Carcinogenesis 29, 244-251.

Gazzerro, E. and Canalis, E. (2006). Bone morphogenetic proteins and theirantagonists. Rev. Endocr. Metab. Disord. 7, 51-65.

Ge, G., Hopkins, D. R., Ho, W. B. and Greenspan, D. S. (2005). GDF11 forms abone morphogenetic protein 1-activated latent complex that can modulatenerve growth factor-induced differentiation of PC12 cells. Mol. Cell. Biol. 25,5846-5858.

Germain, S., Howell, M., Esslemont, G. M. and Hill, C. S. (2000).Homeodomain and winged-helix transcription factors recruit activated Smads todistinct promoter elements via a common Smad interaction motif. Genes Dev.14, 435-451.

Gomis, R. R., Alarcon, C., He, W., Wang, Q., Seoane, J., Lash, A. andMassagué, J. (2006a). A FoxO-Smad synexpression group in humankeratinocytes. Proc. Natl. Acad. Sci. USA 103, 12747-12752.

Gomis, R. R., Alarcon, C., Nadal, C., Van Poznak, C. and Massagué, J.(2006b). C/EBP at the core of the TGF cytostatic response and its evasion inmetastatic breast cancer cells. Cancer Cell 10, 203-214.

Goold, C. P. and Davis, G. W. (2007). The BMP ligand Gbb gates the expressionof synaptic homeostasis independent of synaptic growth control. Neuron 56,109-123.

Gordon, K. J. and Blobe, G. C. (2008). Role of transforming growth factor-superfamily signaling pathways in human disease. Biochim. Biophys. Acta 1782,197-228.

Goto, K., Kamiya, Y., Imamura, T., Miyazono, K. and Miyazawa, K. (2007).Selective inhibitory effects of Smad6 on bone morphogenetic protein type Ireceptors. J. Biol. Chem. 282, 20603-20611.

Goumans, M. J., Valdimarsdottir, G., Itoh, S., Lebrin, F., Larsson, J.,Mummery, C., Karlsson, S. and ten Dijke, P. (2003). Activin receptor-likekinase (ALK)1 is an antagonistic mediator of lateral TGF/ALK5 signaling. Mol.Cell 12, 817-828.

Greber, B., Lehrach, H. and Adjaye, J. (2007). Fibroblast growth factor 2modulates transforming growth factor signaling in mouse embryonicfibroblasts and human ESCs (hESCs) to support hESC self-renewal. Stem Cells25, 455-464.

Gregory, P. A., Bert, A. G., Paterson, E. L., Barry, S. C., Tsykin, A., Farshid, G.,Vadas, M. A., Khew-Goodall, Y. and Goodall, G. J. (2008). The miR-200family and miR-205 regulate epithelial to mesenchymal transition by targetingZEB1 and SIP1. Nat. Cell Biol. 10, 593-601.

Grieder, N. C., Marty, T., Ryoo, H. D., Mann, R. S. and Affolter, M. (1997).Synergistic activation of a Drosophila enhancer by HOM/EXD and DPP signaling.EMBO J. 16, 7402-7410.

Grienenberger, A., Merabet, S., Manak, J., Iltis, I., Fabre, A., Berenger, H.,Scott, M. P., Pradel, J. and Graba, Y. (2003). Tgf signaling acts on a Hoxresponse element to confer specificity and diversity to Hox protein function.Development 130, 5445-5455.

Grinberg, A. V. and Kerppola, T. (2003). Both Max and TFE3 cooperate withSmad proteins to bind the plasminogen activator inhibitor-1 promoter, but theyhave opposite effects on transcriptional activity. J. Biol. Chem. 278, 11227-11236.

Groppe, J., Hinck, C. S., Samavarchi-Tehrani, P., Zubieta, C., Schuermann, J.P., Taylor, A. B., Schwarz, P. M., Wrana, J. L. and Hinck, A. P. (2008).Cooperative assembly of TGF- superfamily signaling complexes is mediated by

two disparate mechanisms and distinct modes of receptor binding. Mol. Cell 29,157-168.

Guo, X., Ramirez, A., Waddell, D. S., Li, Z., Liu, X. and Wang, X. F. (2008a).Axin and GSK3 control Smad3 protein stability and modulate TGF signaling.Genes Dev. 22, 106-120.

Guo, X., Waddell, D. S., Wang, W., Wang, Z., Liberati, N. T., Yong, S., Liu, X.and Wang, X. F. (2008b). Ligand-dependent ubiquitination of Smad3 isregulated by casein kinase 12, an inhibitor of TGF- signaling. Oncogene 27,7235-7247.

Hanai, J., Chen, L. F., Kanno, T., Ohtani-Fujita, N., Kim, W. Y., Guo, W. H.,Imamura, T., Ishidou, Y., Fukuchi, M., Shi, M. J. et al. (1999). Interaction andfunctional cooperation of PEBP2/CBF with Smads. Synergistic induction of theimmunoglobulin germline C promoter. J. Biol. Chem. 274, 31577-31582.

Harrison, C. A., Gray, P. C., Vale, W. W. and Robertson, D. M. (2005).Antagonists of activin signaling: mechanisms and potential biologicalapplications. Trends Endocrinol. Metab. 16, 73-78.

Hata, A., Seoane, J., Lagna, G., Montalvo, E., Hemmati-Brivanlou, A. andMassague, J. (2000). OAZ uses distinct DNA- and protein-binding zinc fingersin separate BMP-Smad and Olf signaling pathways. Cell 100, 229-240.

Hu, H., Milstein, M., Bliss, J. M., Thai, M., Malhotra, G., Huynh, L. C. andColicelli, J. (2008). Integration of transforming growth factor and RASsignaling silences a RAB5 guanine nucleotide exchange factor and enhancesgrowth factor-directed cell migration. Mol. Cell. Biol. 28, 1573-1583.

Hu, M. C. and Rosenblum, N. D. (2005). Smad1, -catenin and Tcf4 associate ina molecular complex with the Myc promoter in dysplastic renal tissue andcooperate to control Myc transcription. Development 132, 215-225.

Hua, X., Miller, Z. A., Wu, G., Shi, Y. and Lodish, H. F. (1999). Specificity intransforming growth factor -induced transcription of the plasminogen activatorinhibitor-1 gene: interactions of promoter DNA, transcription factor E3, andSmad proteins. Proc. Natl. Acad. Sci. USA 96, 13130-13135.

Hua, X., Miller, Z. A., Benchabane, H., Wrana, J. L. and Lodish, H. F. (2000).Synergism between transcription factors TFE3 and Smad3 in transforminggrowth factor-beta-induced transcription of the Smad7 gene. J. Biol. Chem. 275,33205-33208.

Huminiecki, L., Goldovsky, L., Freilich, S., Moustakas, A., Ouzounis, C. andHeldin, C. H. (2009). Emergence, development and diversification of the TGF-signalling pathway within the animal kingdom. BMC Evol. Biol. 9, 28.

Hussein, S. M., Duff, E. K. and Sirard, C. (2003). Smad4 and -catenin co-activators functionally interact with lymphoid-enhancing factor to regulategraded expression of Msx2. J. Biol. Chem. 278, 48805-48814.

Ikushima, H., Komuro, A., Isogaya, K., Shinozaki, M., Hellman, U.,Miyazawa, K. and Miyazono, K. (2008). An Id-like molecule, HHM, is asynexpression group-restricted regulator of TGF- signalling. EMBO J. 27, 2955-2965.

Imoto, S., Ohbayashi, N., Ikeda, O., Kamitani, S., Muromoto, R., Sekine, Y.and Matsuda, T. (2008). Sumoylation of Smad3 stimulates its nuclear exportduring PIASy-mediated suppression of TGF- signaling. Biochem. Biophys. Res.Commun. 370, 359-365.

Ishimura, A., Ng, J. K., Taira, M., Young, S. G. and Osada, S. (2006). Man1, aninner nuclear membrane protein, regulates vascular remodeling by modulatingtransforming growth factor signaling. Development 133, 3919-3928.

Ishimura, A., Chida, S. and Osada, S. (2008). Man1, an inner nuclear membraneprotein, regulates left-right axis formation by controlling nodal signaling in anode-independent manner. Dev. Dyn. 237, 3565-3576.

Itoh, F., Itoh, S., Goumans, M. J., Valdimarsdottir, G., Iso, T., Dotto, G. P.,Hamamori, Y., Kedes, L., Kato, M. and ten Dijke, P. (2004). Synergy andantagonism between Notch and BMP receptor signaling pathways in endothelialcells. EMBO J. 23, 541-551.

Itoh, S. and ten Dijke, P. (2007). Negative regulation of TGF- receptor/Smadsignal transduction. Curr. Opin. Cell Biol. 19, 176-184.

Jiang, X., Xia, L., Chen, D., Yang, Y., Huang, H., Yang, L., Zhao, Q., Shen, L.and Wang, J. (2008). Otefin, a nuclear membrane protein, determines the fateof germline stem cells in Drosophila via interaction with Smad complexes. Dev.Cell 14, 494-506.

Jin, Q., Ding, W. and Mulder, K. M. (2007). Requirement for the dynein lightchain km23-1 in a Smad2-dependent transforming growth factor- signalingpathway. J. Biol. Chem. 282, 19122-19132.

Jin, W., Takagi, T., Kanesashi, S. N., Kurahashi, T., Nomura, T., Harada, J. andIshii, S. (2006). Schnurri-2 controls BMP-dependent adipogenesis via interactionwith Smad proteins. Dev. Cell 10, 461-471.

Kahlem, P. and Newfeld, S. J. (2009). Informatics approaches to understandingTGF pathway regulation. Development 136, 3729-3740.

Kajino, T., Omori, E., Ishii, S., Matsumoto, K. and Ninomiya-Tsuji, J. (2007).TAK1 MAPK kinase kinase mediates transforming growth factor- signaling bytargeting SnoN oncoprotein for degradation. J. Biol. Chem. 282, 9475-9481.

Kalkan, T., Iwasaki, Y., Park, C. Y. and Thomsen, G. H. (2009). TRAF4 is apositive regulator of TGF signaling that affects neural crest formation. Mol.Biol. Cell 20, 3436-3450.

Kalo, E., Buganim, Y., Shapira, K. E., Besserglick, H., Goldfinger, N., Weisz,L., Stambolsky, P., Henis, Y. I. and Rotter, V. (2007). Mutant p53 attenuates D

EVELO

PMENT

Page 14: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3712

the SMAD-dependent transforming growth factor 1 (TGF-1) signalingpathway by repressing the expression of TGF- receptor type II. Mol. Cell. Biol.27, 8228-8242.

Kamiya, Y., Miyazono, K. and Miyazawa, K. (2008). Specificity of the inhibitoryeffects of Dad on TGF- family type I receptors, Thickveins, Saxophone, andBaboon in Drosophila. FEBS Lett. 582, 2496-2500.

Kang, J. S., Saunier, E. F., Akhurst, R. J. and Derynck, R. (2008). The type I TGF- receptor is covalently modified and regulated by sumoylation. Nat. Cell Biol.10, 654-664.

Kang, Y., Chen, C. R. and Massagué, J. (2003). A self-enabling TGF responsecoupled to stress signaling: Smad engages stress response factor ATF3 for Id1repression in epithelial cells. Mol. Cell 11, 915-926.

Karaulanov, E., Knochel, W. and Niehrs, C. (2004). Transcriptional regulation ofBMP4 synexpression in transgenic Xenopus. EMBO J. 23, 844-856.

Kardassis, D., Murphy, C., Fotsis, T., Moustakas, A. and Stournaras, C.(2009). Control of transforming growth factor signal transduction by smallGTPases. FEBS J. 276, 2947-2965.

Katagiri, T., Imada, M., Yanai, T., Suda, T., Takahashi, N. and Kamijo, R.(2002). Identification of a BMP-responsive element in Id1, the gene for inhibitionof myogenesis. Genes Cells 7, 949-960.

Kato, Y., Habas, R., Katsuyama, Y., Näär, A. M. and He, X. (2002). Acomponent of the ARC/Mediator complex required for TGF/Nodal signalling.Nature 418, 641-646.

Koinuma, D., Shinozaki, M., Komuro, A., Goto, K., Saitoh, M., Hanyu, A.,Ebina, M., Nukiwa, T., Miyazawa, K., Imamura, T. et al. (2003). Arkadiaamplifies TGF- superfamily signalling through degradation of Smad7. EMBO J.22, 6458-6470.

Koinuma, D., Tsutsumi, S., Kamimura, N., Taniguchi, H., Miyazawa, K.,Sunamura, M., Imamura, T., Miyazono, K. and Aburatani, H. (2009).Chromatin immunoprecipitation on microarray analysis of Smad2/3 binding sitesreveals roles of ETS1 and TFAP2A in transforming growth factor signaling.Mol. Cell. Biol. 29, 172-186.

Korchynskyi, O. and ten Dijke, P. (2002). Identification and functionalcharacterization of distinct critically important bone morphogenetic protein-specific response elements in the Id1 promoter. J. Biol. Chem. 277, 4883-4891.

Korpal, M., Lee, E. S., Hu, G. and Kang, Y. (2008). The miR-200 family inhibitsepithelial-mesenchymal transition and cancer cell migration by direct targetingof E-cadherin transcriptional repressors ZEB1 and ZEB2. J. Biol. Chem. 283,14910-14914.

Kowanetz, M., Valcourt, U., Bergström, R., Heldin, C. H. and Moustakas, A.(2004). Id2 and Id3 define the potency of cell proliferation and differentiationresponses to tranforming growth factor and bone morphogenetic protein.Mol. Cell. Biol. 24, 4241-4254.

Kowanetz, M., Lönn, P., Vanlandewijck, M., Kowanetz, K., Heldin, C. H. andMoustakas, A. (2008). TGF induces SIK to negatively regulate type I receptorkinase signaling. J. Cell Biol. 182, 655-662.

Ku, M., Howard, S., Ni, W., Lagna, G. and Hata, A. (2006). OAZ regulates bonemorphogenetic protein signaling through Smad6 activation. J. Biol. Chem. 281,5277-5287.

Labbé, E., Letamendia, A. and Attisano, L. (2000). Association of Smads withlymphoid enhancer binding factor 1/T cell- specific factor mediates cooperativesignaling by the transforming growth factor- and wnt pathways. Proc. Natl.Acad. Sci. USA 97, 8358-8363.

Le Good, J. A., Joubin, K., Giraldez, A. J., Ben-Haim, N., Beck, S., Chen, Y.,Schier, A. F. and Constam, D. B. (2005). Nodal stability determines signalingrange. Curr. Biol. 15, 31-36.

Le Scolan, E., Zhu, Q., Wang, L., Bandyopadhyay, A., Javelaud, D., Mauviel,A., Sun, L. and Luo, K. (2008). Transforming growth factor- suppresses theability of Ski to inhibit tumor metastasis by inducing its degradation. Cancer Res.68, 3277-3285.

Lee, H. H. and Frasch, M. (2005). Nuclear integration of positive Dpp signals,antagonistic Wg inputs and mesodermal competence factors during Drosophilavisceral mesoderm induction. Development 132, 1429-1442.

Lee, H. J., Yun, C. H., Lim, S. H., Kim, B. C., Baik, K. G., Kim, J. M., Kim, W. H.and Kim, S. J. (2007). SRF is a nuclear repressor of Smad3-mediated TGF-signaling. Oncogene 26, 173-185.

Lee, M. K., Pardoux, C., Hall, M. C., Lee, P. S., Warburton, D., Qing, J., Smith,S. M. and Derynck, R. (2007). TGF- activates Erk MAP kinase signallingthrough direct phosphorylation of ShcA. EMBO J. 26, 3957-3967.

Lee-Hoeflich, S. T., Zhao, X., Mehra, A. and Attisano, L. (2005). The Drosophilatype II receptor, Wishful thinking, binds BMP and myoglianin to activate multipleTGF family signaling pathways. FEBS Lett. 579, 4615-4621.

Lei, S., Dubeykovskiy, A., Chakladar, A., Wojtukiewicz, L. and Wang, T. C.(2004). The murine gastrin promoter is synergistically activated by transforminggrowth factor-/Smad and Wnt signaling pathways. J. Biol. Chem. 279, 42492-42502.

Levy, L., Howell, M., Das, D., Harkin, S., Episkopou, V. and Hill, C. S. (2007).Arkadia activates Smad3/Smad4-dependent transcription by triggering signal-induced SnoN degradation. Mol. Cell. Biol. 27, 6068-6083.

Li, C., Zhu, N. L., Tan, R. C., Ballard, P. L., Derynck, R. and Minoo, P. (2002).

Transforming growth factor- inhibits pulmonary surfactant protein B genetranscription through SMAD3 interactions with NKX2.1 and HNF-3 transcriptionfactors. J. Biol. Chem. 277, 38399-38408.

Li, Z., Hassan, M. Q., Volinia, S., van Wijnen, A. J., Stein, J. L., Croce, C. M.,Lian, J. B. and Stein, G. S. (2008). A microRNA signature for a BMP2-inducedosteoblast lineage commitment program. Proc. Natl. Acad. Sci. USA 105,13906-13911.

Liang, J., Lints, R., Foehr, M. L., Tokarz, R., Yu, L., Emmons, S. W., Liu, J. andSavage-Dunn, C. (2003). The Caenorhabditis elegans schnurri homolog sma-9mediates stage- and cell type-specific responses to DBL-1 BMP-related signaling.Development 130, 6453-6464.

Liang, J., Yu, L., Yin, J. and Savage-Dunn, C. (2007). Transcriptional repressorand activator activities of SMA-9 contribute differentially to BMP-relatedsignaling outputs. Dev. Biol. 305, 714-725.

Lin, F., Morrison, J. M., Wu, W. and Worman, H. J. (2005). MAN1, an integralprotein of the inner nuclear membrane, binds Smad2 and Smad3 andantagonizes transforming growth factor- signaling. Hum. Mol. Genet. 14, 437-445.

Lin, H. K., Bergmann, S. and Pandolfi, P. P. (2004). Cytoplasmic PML function inTGF- signalling. Nature 431, 205-211.

Liu, D., Kang, J. S. and Derynck, R. (2004). TGF--activated Smad3 repressesMEF2-dependent transcription in myogenic differentiation. EMBO J. 23, 1557-1566.

Liu, I. M., Schilling, S. H., Knouse, K. A., Choy, L., Derynck, R. and Wang, X.F. (2008). TGF-stimulated Smad1/5 phosphorylation requires the ALK5 L45 loopand mediates the pro-migratory TGF switch. EMBO J. 28, 88-98.

Lönn, P., Morén, A., Raja, E., Dahl, M. and Moustakas, A. (2009). Regulatingthe stability of TGF receptors and Smads. Cell Res. 19, 21-35.

Lopez-Rovira, T., Chalaux, E., Rosa, J. L., Bartrons, R. and Ventura, F. (2000).Interaction and functional cooperation of NF-B with Smads. Transcriptionalregulation of the junB promoter. J. Biol. Chem. 275, 28937-28946.

Lopez-Rovira, T., Chalaux, E., Massagué, J., Rosa, J. L. and Ventura, F. (2002).Direct binding of Smad1 and Smad4 to two distinct motifs mediates bonemorphogenetic protein-specific transcriptional activation of Id1 gene. J. Biol.Chem. 277, 3176-3185.

Maduzia, L. L., Roberts, A. F., Wang, H., Lin, X., Chin, L. J., Zimmerman, C.M., Cohen, S., Feng, X. H. and Padgett, R. W. (2005). C. elegans serine-threonine kinase KIN-29 modulates TGF signaling and regulates body sizeformation. BMC Dev. Biol. 5, 8.

Major, M. B. and Jones, D. A. (2004). Identification of a gadd45 3� enhancerthat mediates SMAD3- and SMAD4-dependent transcriptional induction bytransforming growth factor . J. Biol. Chem. 279, 5278-5287.

Marinari, B., Moretti, F., Botti, E., Giustizieri, M. L., Descargues, P., Giunta,A., Stolfi, C., Ballaro, C., Papoutsaki, M., Alema, S. et al. (2008). The tumorsuppressor activity of IKK in stratified epithelia is exerted in part via the TGF-antiproliferative pathway. Proc. Natl. Acad. Sci. USA 105, 17091-17096.

Martello, G., Zacchigna, L., Inui, M., Montagner, M., Adorno, M., Mamidi,A., Morsut, L., Soligo, S., Tran, U., Dupont, S. et al. (2007). MicroRNAcontrol of Nodal signalling. Nature 449, 183-188.

Marty, T., Muller, B., Basler, K. and Affolter, M. (2000). Schnurri mediates Dpp-dependent repression of brinker transcription. Nat. Cell Biol. 2, 745-749.

Massagué, J. (2008). TGF in cancer. Cell 134, 215-230.Massagué, J., Seoane, J. and Wotton, D. (2005). Smad transcription factors.

Genes Dev. 19, 2783-2810.Mavrakis, K. J., Andrew, R. L., Lee, K. L., Petropoulou, C., Dixon, J. E.,

Navaratnam, N., Norris, D. P. and Episkopou, V. (2007). Arkadia enhancesNodal/TGF- signaling by coupling phospho-Smad2/3 activity and turnover. PLoSBiol. 5, e67.

Messenger, N. J., Kabitschke, C., Andrews, R., Grimmer, D., Nunez Miguel,R., Blundell, T. L., Smith, J. C. and Wardle, F. C. (2005). Functional specificityof the Xenopus T-domain protein Brachyury is conferred by its ability to interactwith Smad1. Dev. Cell 8, 599-610.

Miles, W. O., Jaffray, E., Campbell, S. G., Takeda, S., Bayston, L. J., Basu, S.P., Li, M., Raftery, L. A., Ashe, M. P., Hay, R. T. et al. (2008). MedeaSUMOylation restricts the signaling range of the Dpp morphogen in theDrosophila embryo. Genes Dev. 22, 2578-2590.

Mochizuki, T., Miyazaki, H., Hara, T., Furuya, T., Imamura, T., Watabe, T. andMiyazono, K. (2004). Roles for the MH2 domain of Smad7 in the specificinhibition of transforming growth factor- superfamily signaling. J. Biol. Chem.279, 31568-31574.

Morén, A., Hellman, U., Inada, Y., Imamura, T., Heldin, C. H. and Moustakas,A. (2003). Differential ubiquitination defines the functional status of the tumorsuppressor Smad4. J. Biol. Chem. 278, 33571-33582.

Moustakas, A. and Heldin, C. H. (2005). Non-Smad TGF- signals. J. Cell Sci.118, 3573-3584.

Moustakas, A. and Heldin, C. H. (2007). Signaling networks guiding epithelial-mesenchymal transitions during embryogenesis and cancer progression. CancerSci. 98, 1512-1520.

Moustakas, A. and Heldin, C. H. (2008). Dynamic control of TGF- signaling andits links to the cytoskeleton. FEBS Lett. 582, 2051-2065.

REVIEW Development 136 (22)

DEVELO

PMENT

Page 15: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3713REVIEWDevelopment 136 (22)

Nagano, Y., Mavrakis, K. J., Lee, K. L., Fujii, T., Koinuma, D., Sase, H., Yuki,K., Isogaya, K., Saitoh, M., Imamura, T. et al. (2007). Arkadia inducesdegradation of SnoN and c-Ski to enhance transforming growth factor-signaling. J. Biol. Chem. 282, 20492-20501.

Nakashima, A., Katagiri, T. and Tamura, M. (2005). Cross-talk between Wntand bone morphogenetic protein 2 (BMP-2) signaling in differentiation pathwayof C2C12 myoblasts. J. Biol. Chem. 280, 37660-37668.

Nguyen, M., Parker, L. and Arora, K. (2000). Identification of maverick, a novelmember of the TGF- superfamily in Drosophila. Mech. Dev. 95, 201-206.

Niehrs, C. and Pollet, N. (1999). Synexpression groups in eukaryotes. Nature 402,483-487.

Niimi, H., Pardali, K., Vanlandewijck, M., Heldin, C. H. and Moustakas, A.(2007). Notch signaling is necessary for epithelial growth arrest by TGF-. J. CellBiol. 176, 695-707.

Nishita, M., Hashimoto, M. K., Ogata, S., Laurent, M. N., Ueno, N., Shibuya,H. and Cho, K. W. (2000). Interaction between Wnt and TGF- signallingpathways during formation of Spemann’s organizer. Nature 403, 781-785.

Ogawa, K., Saito, A., Matsui, H., Suzuki, H., Ohtsuka, S., Shimosato, D.,Morishita, Y., Watabe, T., Niwa, H. and Miyazono, K. (2007). Activin-Nodalsignaling is involved in propagation of mouse embryonic stem cells. J. Cell Sci.120, 55-65.

Onuma, Y., Takahashi, S., Yokota, C. and Asashima, M. (2002). Multiple nodal-related genes act coordinately in Xenopus embryogenesis. Dev. Biol. 241, 94-105.

Osada, S., Ohmori, S. Y. and Taira, M. (2003). XMAN1, an inner nuclearmembrane protein, antagonizes BMP signaling by interacting with Smad1 inXenopus embryos. Development 130, 1783-1794.

Osada, S. I., Saijoh, Y., Frisch, A., Yeo, C. Y., Adachi, H., Watanabe, M.,Whitman, M., Hamada, H. and Wright, C. V. (2000). Activin/nodalresponsiveness and asymmetric expression of a Xenopus nodal-related geneconverge on a FAST-regulated module in intron 1. Development 127, 2503-2514.

Ozdamar, B., Bose, R., Barrios-Rodiles, M., Wang, H. R., Zhang, Y. andWrana, J. L. (2005). Regulation of the polarity protein Par6 by TGF receptorscontrols epithelial cell plasticity. Science 307, 1603-1609.

Pan, D., Estevez-Salmeron, L. D., Stroschein, S. L., Zhu, X., He, J., Zhou, S.and Luo, K. (2005). The integral inner nuclear membrane protein MAN1physically interacts with the R-Smad proteins to repress signaling by thetransforming growth factor- superfamily of cytokines. J. Biol. Chem. 280,15992-16001.

Pardali, E., Xie, X. Q., Tsapogas, P., Itoh, S., Arvanitidis, K., Heldin, C. H., tenDijke, P., Grundstrom, T. and Sideras, P. (2000). Smad and AML proteinssynergistically confer transforming growth factor 1 responsiveness to humangerm-line IgA genes. J. Biol. Chem. 275, 3552-3560.

Pardali, K., Kurisaki, A., Morén, A., ten Dijke, P., Kardassis, D. andMoustakas, A. (2000). Role of Smad proteins and transcription factor Sp1 inp21Waf1/Cip1 regulation by transforming growth factor-. J. Biol. Chem. 275,29244-29256.

Patterson, G. I. and Padgett, R. W. (2000). TGF -related pathways. Roles inCaenorhabditis elegans development. Trends Genet. 16, 27-33.

Patterson, G. I., Koweek, A., Wong, A., Liu, Y. and Ruvkun, G. (1997). TheDAF-3 Smad protein antagonizes TGF--related receptor signaling in theCaenorhabditis elegans dauer pathway. Genes Dev. 11, 2679-2690.

Peñuelas, S., Anido, J., Prieto-Sánchez, R. M., Folch, G., Barba, I., Cuartas, I.,García-Dorado, D., Poca, M. A., Sahuquillo, J., Baselga, J. et al. (2009).TGF- increases glioma-initiating cell self-renewal through the induction of LIF inhuman glioblastoma. Cancer Cell 15, 315-327.

Pyrowolakis, G., Hartmann, B., Muller, B., Basler, K. and Affolter, M. (2004).A simple molecular complex mediates widespread BMP-induced repressionduring Drosophila development. Dev. Cell 7, 229-240.

Qiu, P., Feng, X. H. and Li, L. (2003). Interaction of Smad3 and SRF-associatedcomplex mediates TGF-1 signals to regulate SM22 transcription duringmyofibroblast differentiation. J. Mol. Cell. Cardiol. 35, 1407-1420.

Raju, G. P., Dimova, N., Klein, P. S. and Huang, H. C. (2003). SANE, a novel LEMdomain protein, regulates bone morphogenetic protein signaling throughinteraction with Smad1. J. Biol. Chem. 278, 428-437.

Ramis, J. M., Collart, C. and Smith, J. C. (2007). Xnrs and activin regulatedistinct genes during Xenopus development: activin regulates cell division. PLoSONE 2, e213.

Rifkin, D. B. (2005). Latent transforming growth factor- (TGF-) binding proteins:orchestrators of TGF- availability. J. Biol. Chem. 280, 7409-7412.

Rossi, D. J., Jamieson, C. H. and Weissman, I. L. (2008). Stems cells and thepathways to aging and cancer. Cell 132, 681-696.

Saijoh, Y., Adachi, H., Mochida, K., Ohishi, S., Hirao, A. and Hamada, H.(1999). Distinct transcriptional regulatory mechanisms underlie left-rightasymmetric expression of lefty-1 and lefty-2. Genes Dev. 13, 259-269.

Saka, Y., Hagemann, A. I., Piepenburg, O. and Smith, J. C. (2007). Nuclearaccumulation of Smad complexes occurs only after the midblastula transition inXenopus. Development 134, 4209-4218.

Sanchez-Elsner, T., Botella, L. M., Velasco, B., Corbi, A., Attisano, L. and

Bernabeu, C. (2001). Synergistic cooperation between hypoxia andtransforming growth factor- pathways on human vascular endothelial growthfactor gene expression. J. Biol. Chem. 276, 38527-38535.

Sapkota, G., Alarcon, C., Spagnoli, F. M., Brivanlou, A. H. and Massagué, J.(2007). Balancing BMP signaling through integrated inputs into the Smad1linker. Mol. Cell 25, 441-454.

Sasai, N., Yakura, R., Kamiya, D., Nakazawa, Y. and Sasai, Y. (2008).Ectodermal factor restricts mesoderm differentiation by inhibiting p53. Cell 133,878-890.

Savage-Dunn, C. (2005). TGF- signaling. WormBook, 1-12.http://www.wormbook.org/.

Savage-Dunn, C., Maduzia, L. L., Zimmerman, C. M., Roberts, A. F., Cohen,S., Tokarz, R. and Padgett, R. W. (2003). Genetic screen for small body sizemutants in C. elegans reveals many TGF pathway components. Genesis 35,239-247.

Schmierer, B. and Hill, C. S. (2007). TGF-SMAD signal transduction: molecularspecificity and functional flexibility. Nat. Rev. Mol. Cell. Biol. 8, 970-982.

Schmierer, B., Tournier, A. L., Bates, P. A. and Hill, C. S. (2008). Mathematicalmodeling identifies Smad nucleocytoplasmic shuttling as a dynamic signal-interpreting system. Proc. Natl. Acad. Sci. USA 105, 6608-6613.

Seoane, J., Pouponnot, C., Staller, P., Schader, M., Eilers, M. and Massagué,J. (2001). TGF influences Myc, Miz-1 and Smad to control the CDK inhibitorp15INK4b. Nat. Cell Biol. 3, 400-408.

Seoane, J., Le H. V., Shen, L., Anderson, S. A. and Massagué, J. (2004).Integration of Smad and forkhead pathways in the control of neuroepithelialand glioblastoma cell proliferation. Cell 117, 211-223.

Serpe, M., Ralston, A., Blair, S. S. and O’Connor, M. B. (2005). Matchingcatalytic activity to developmental function: tolloid-related processes Sog inorder to help specify the posterior crossvein in the Drosophila wing.Development 132, 2645-2656.

Shakéd, M., Weissmüller, K., Svoboda, H., Hortschansky, P., Nishino, N.,Wolfl, S. and Tucker, K. L. (2008). Histone deacetylases control neurogenesis inembryonic brain by inhibition of BMP2/4 signaling. PLoS ONE 3, e2668.

Shen, J., Dorner, C., Bahlo, A. and Pflügfelder, G. O. (2008). optomotor-blindsuppresses instability at the A/P compartment boundary of the Drosophila wing.Mech. Dev. 125, 233-246.

Shi, W., Chang, C., Nie, S., Xie, S., Wan, M. and Cao, X. (2007). Endofin acts asa Smad anchor for receptor activation in BMP signaling. J. Cell Sci. 120, 1216-1224.

Shimizu, K., Bourillot, P. Y., Nielsen, S. J., Zorn, A. M. and Gurdon, J. B.(2001). Swift is a novel BRCT domain coactivator of Smad2 in transforminggrowth factor signaling. Mol. Cell. Biol. 21, 3901-3912.

Shiratori, H., Sakuma, R., Watanabe, M., Hashiguchi, H., Mochida, K., Sakai,Y., Nishino, J., Saijoh, Y., Whitman, M. and Hamada, H. (2001). Two-stepregulation of left-right asymmetric expression of Pitx2: initiation by nodalsignaling and maintenance by Nkx2. Mol. Cell 7, 137-149.

Simonsson, M., Kanduri, M., Grönroos, E., Heldin, C.-H. and Ericsson, J.(2006). The DNA binding activities of Smad2 and Smad3 are regulated bycoactivator-mediated acetylation. J. Biol. Chem. 281, 39870-39880.

Sivasankaran, R., Vigano, M. A., Muller, B., Affolter, M. and Basler, K. (2000).Direct transcriptional control of the Dpp target omb by the DNA binding proteinBrinker. EMBO J. 19, 6162-6172.

Smith, J. C. and Gurdon, J. B. (2004). Many ways to make a gradient. BioEssays26, 705-706.

Sorrentino, A., Thakur, N., Grimsby, S., Marcusson, A., von Bulow, V.,Schuster, N., Zhang, S., Heldin, C. H. and Landström, M. (2008). The type ITGF- receptor engages TRAF6 to activate TAK1 in a receptor kinase-independent manner. Nat. Cell Biol. 10, 1199-1207.

Sowa, H., Kaji, H., Hendy, G. N., Canaff, L., Komori, T., Sugimoto, T. andChihara, K. (2004). Menin is required for bone morphogenetic protein 2- andtransforming growth factor -regulated osteoblastic differentiation throughinteraction with Smads and Runx2. J. Biol. Chem. 279, 40267-40275.

Su, Y., Zhang, L., Gao, X., Meng, F., Wen, J., Zhou, H., Meng, A. and Chen, Y.G. (2007). The evolutionally conserved activity of Dapper2 in antagonizing TGF-signaling. FASEB J. 21, 682-690.

Sun, Q., Zhang, Y., Yang, G., Chen, X., Cao, G., Wang, J., Sun, Y., Zhang, P.,Fan, M., Shao, N. et al. (2008). Transforming growth factor--regulated miR-24promotes skeletal muscle differentiation. Nucleic Acids Res. 36, 2690-2699.

Suzuki, A., Raya, A., Kawakami, Y., Morita, M., Matsui, T., Nakashima, K.,Gage, F. H., Rodriguez-Esteban, C. and Izpisua Belmonte, J. C. (2006).Nanog binds to Smad1 and blocks bone morphogenetic protein-induceddifferentiation of embryonic stem cells. Proc. Natl. Acad. Sci. USA 103, 10294-10299.

Takebayashi-Suzuki, K., Funami, J., Tokumori, D., Saito, A., Watabe, T.,Miyazono, K., Kanda, A. and Suzuki, A. (2003). Interplay between the tumorsuppressor p53 and TGF signaling shapes embryonic body axes in Xenopus.Development 130, 3929-3939.

ten Dijke, P. and Heldin, C. H. (2006). The Smad family. In Smad SignalTransduction, vol. 5 (ed. P. ten Dijke and C. H. Heldin), pp. 1-13. Dordrecht, TheNetherlands: Springer. D

EVELO

PMENT

Page 16: The regulation of TGF signal transduction · signal transduction in metazoans and emphasize events crucial for embryonic development. Introduction The human transforming growth factor

3714

ten Dijke, P. and Arthur, H. M. (2007). Extracellular control of TGF signalling invascular development and disease. Nat. Rev. Mol. Cell Biol. 8, 857-869.

Thuault, S., Tan, E. J., Peinado, H., Cano, A., Heldin, C. H. and Moustakas, A.(2008). HMGA2 and Smads co-regulate SNAIL1 expression during induction ofepithelial-to-mesenchymal transition. J. Biol. Chem. 283, 33437-33446.

Tsukazaki, T., Chiang, T. A., Davison, A. F., Attisano, L. and Wrana, J. L.(1998). SARA, a FYVE domain protein that recruits Smad2 to the TGF receptor.Cell 95, 779-791.

Tsuneizumi, K., Nakayama, T., Kamoshida, Y., Kornberg, T. B., Christian, J. L.and Tabata, T. (1997). Daughters against dpp modulates dpp organizingactivity in Drosophila wing development. Nature 389, 627-631.

Tu, A. W. and Luo, K. (2007). Acetylation of Smad2 by the co-activator p300regulates activin and transforming growth factor response. J. Biol. Chem. 282,21187-21196.

Umulis, D., O’Connor, M. B. and Blair, S. S. (2009). The extracellular regulationof bone morphogenetic protein signaling. Development 136, 3715-3728.

Valcourt, U., Thuault, S., Pardali, K., Heldin, C. H. and Moustakas, A. (2007).Functional role of Meox2 during the epithelial cytostatic response to TGF-. Mol.Oncol. 1, 55-71.

Van Beek, J. P., Kennedy, L., Rockel, J. S., Bernier, S. M. and Leask, A. (2006).The induction of CCN2 by TGF1 involves Ets-1. Arthritis Res. Ther. 8, R36.

Varelas, X., Sakuma, R., Samavarchi-Tehrani, P., Peerani, R., Rao, B. M.,Dembowy, J., Yaffe, M. B., Zandstra, P. W. and Wrana, J. L. (2008). TAZcontrols Smad nucleocytoplasmic shuttling and regulates human embryonicstem-cell self-renewal. Nat. Cell Biol. 10, 837-848.

Wang, B., Suzuki, H. and Kato, M. (2008). Roles of mono-ubiquitinated Smad4in the formation of Smad transcriptional complexes. Biochem. Biophys. Res.Commun. 376, 288-292.

Wang, G., Matsuura, I., He, D. and Liu, F. (2009). Transforming growth factor--inducible phosphorylation of Smad3. J. Biol. Chem. 284, 9663-9673.

Wang, Q., Wei, X., Zhu, T., Zhang, M., Shen, R., Xing, L., O’Keefe, R. J. andChen, D. (2007). Bone morphogenetic protein 2 activates Smad6 genetranscription through bone-specific transcription factor Runx2. J. Biol. Chem.282, 10742-10748.

Watabe, T. and Miyazono, K. (2009). Roles of TGF- family signaling in stem cellrenewal and differentiation. Cell Res. 19, 103-115.

Watanabe, M., Rebbert, M. L., Andreazzoli, M., Takahashi, N., Toyama, R.,Zimmerman, S., Whitman, M. and Dawid, I. B. (2002). Regulation of theLim-1 gene is mediated through conserved FAST-1/FoxH1 sites in the first intron.Dev. Dyn. 225, 448-456.

Wharton, K. and Derynck, R. (2009). TGF family signaling: novel insights indevelopment and disease. Development 136, 3691-3697.

Wilkinson, D. S., Tsai, W. W., Schumacher, M. A. and Barton, M. C. (2008).Chromatin-bound p53 anchors activated Smads and the mSin3A corepressor toconfer transforming-growth-factor--mediated transcription repression. Mol.Cell. Biol. 28, 1988-1998.

Wolfman, N. M., McPherron, A. C., Pappano, W. N., Davies, M. V., Song, K.,Tomkinson, K. N., Wright, J. F., Zhao, L., Sebald, S. M., Greenspan, D. S. etal. (2003). Activation of latent myostatin by the BMP-1/tolloid family ofmetalloproteinases. Proc. Natl. Acad. Sci. USA 100, 15842-15846.

Wrana, J. L., Ozdamar, B., Le Roy, C. and Benchabane, H. (2008). Signalingreceptors of the TGF- family. In The TGF- Family (ed. R. Derynck and K.Miyazono), pp. 179-202. Cold Spring Harbor, NY: Cold Spring Harbor LaboratoryPress.

Wrighton, K. H., Lin, X. and Feng, X. H. (2008). Critical regulation of TGFsignaling by Hsp90. Proc. Natl. Acad. Sci. USA 105, 9244-9249.

Wrighton, K. H., Lin, X. and Feng, X. H. (2009). Phospho-control of TGF-superfamily signaling. Cell Res. 19, 8-20.

Xi, Q., He, W., Zhang, X. H., Le H. V. and Massagué, J. (2008). Genome-wideimpact of the BRG1 SWI/SNF chromatin remodeler on the transforming growthfactor transcriptional program. J. Biol. Chem. 283, 1146-1155.

Xu, L., Yao, X., Chen, X., Lu, P., Zhang, B. and Ip, Y. T. (2007). Msk is requiredfor nuclear import of TGF-/BMP-activated Smads. J. Cell Biol. 178, 981-994.

Xu, M., Kirov, N. and Rushlow, C. (2005). Peak levels of BMP in the Drosophilaembryo control target genes by a feed-forward mechanism. Development 132,1637-1647.

Xu, R. H., Sampsell-Barron, T. L., Gu, F., Root, S., Peck, R. M., Pan, G., Yu, J.,Antosiewicz-Bourget, J., Tian, S., Stewart, R. et al. (2008). NANOG is adirect target of TGF/activin-mediated SMAD signaling in human ESCs. CellStem Cell 3, 196-206.

Xu, X., Yin, Z., Hudson, J. B., Ferguson, E. L. and Frasch, M. (1998). Smadproteins act in combination with synergistic and antagonistic regulators to targetDpp responses to the Drosophila mesoderm. Genes Dev. 12, 2354-2370.

Xu, X., Han, J., Ito, Y., Bringas, P., Jr, Deng, C. and Chai, Y. (2008).Ectodermal Smad4 and p38 MAPK are functionally redundant in mediatingTGF-/BMP signaling during tooth and palate development. Dev. Cell 15, 322-329.

Yamashita, M., Fatyol, K., Jin, C., Wang, X., Liu, Z. and Zhang, Y. E. (2008).TRAF6 mediates Smad-independent activation of JNK and p38 by TGF-. Mol.Cell 31, 918-924.

Yang, G., Li, Y., Nishimura, E. K., Xin, H., Zhou, A., Guo, Y., Dong, L.,Denning, M. F., Nickoloff, B. J. and Cui, R. (2008). Inhibition of PAX3 by TGF- modulates melanocyte viability. Mol. Cell 32, 554-563.

Yang, L. and Moses, H. L. (2008). Transforming growth factor : tumorsuppressor or promoter? Are host immune cells the answer? Cancer Res. 68,9107-9111.

Yano, T., Ito, K., Fukamachi, H., Chi, X. Z., Wee, H. J., Inoue, K., Ida, H.,Bouillet, P., Strasser, A., Bae, S. C. et al. (2006). The RUNX3 tumor suppressorupregulates Bim in gastric epithelial cells undergoing transforming growth factor-induced apoptosis. Mol. Cell. Biol. 26, 4474-4488.

Yao, X., Chen, X., Cottonham, C. and Xu, L. (2008). Preferential utilization ofImp7/8 in nuclear import of Smads. J. Biol. Chem. 283, 22867-22874.

Yoo, J., Ghiassi, M., Jirmanova, L., Balliet, A. G., Hoffman, B., Fornace, A. J.,Jr, Liebermann, D. A., Böttinger, E. P. and Roberts, A. B. (2003).Transforming growth factor--induced apoptosis is mediated by Smad-dependent expression of GADD45b through p38 activation. J. Biol. Chem. 278,43001-43007.

Zavadil, J., Cermak, L., Soto-Nieves, N. and Böttinger, E. P. (2004). Integrationof TGF-/Smad and Jagged1/Notch signalling in epithelial-to-mesenchymaltransition. EMBO J. 23, 1155-1165.

Zavadil, J., Narasimhan, M., Blumenberg, M. and Schneider, R. J. (2007).Transforming growth factor- and microRNA:mRNA regulatory networks inepithelial plasticity. Cells Tissues Organs 185, 157-161.

Zhang, S., Fei, T., Zhang, L., Zhang, R., Chen, F., Ning, Y., Han, Y., Feng, X. H.,Meng, A. and Chen, Y. G. (2007). Smad7 antagonizes transforming growthfactor signaling in the nucleus by interfering with functional Smad-DNAcomplex formation. Mol. Cell. Biol. 27, 4488-4499.

Zhang, Y. and Derynck, R. (2000). Transcriptional regulation of the transforminggrowth factor--inducible mouse germ line Ig constant region gene byfunctional cooperation of Smad, CREB, and AML family members. J. Biol. Chem.275, 16979-16985.

Zhang, Y. W., Yasui, N., Ito, K., Huang, G., Fujii, M., Hanai, J., Nogami, H.,Ochi, T., Miyazono, K. and Ito, Y. (2000). A RUNX2/PEBP2A/CBFA1 mutationdisplaying impaired transactivation and Smad interaction in cleidocranialdysplasia. Proc. Natl. Acad. Sci. USA 97, 10549-10554.

Zheng, X., Wang, J., Haerry, T. E., Wu, A. Y., Martin, J., O’Connor, M. B., Lee,C. H. and Lee, T. (2003). TGF- signaling activates steroid hormone receptorexpression during neuronal remodeling in the Drosophila brain. Cell 112, 303-315.

Zhu, C. C., Boone, J. Q., Jensen, P. A., Hanna, S., Podemski, L., Locke, J., Doe,C. Q. and O’Connor, M. B. (2008). Drosophila Activin- and the Activin-likeproduct Dawdle function redundantly to regulate proliferation in the larval brain.Development 135, 513-521.

Zhu, S., Wang, W., Clarke, D. C. and Liu, X. (2007). Activation of Mps1promotes transforming growth factor--independent Smad signaling. J. Biol.Chem. 282, 18327-18338.

REVIEW Development 136 (22)

DEVELO

PMENT